content
stringlengths
1
15.9M
\section{Introduction} Strongly interacting bosonic systems have attracted a lot of recent interest\cite{Fishers,Scalettar,Rokhsar,Girvin}. Physical realizations include short correlation-length superconductors, granular superconductors, Josephson arrays, the dynamics of flux lattices in type II superconductors, and critical behavior of $^4{\rm He}$ in porous media. The bosonic systems are either tightly bound composites of fermions that act like effective bosonic particles with soft cores, or correspond to bosonic excitations that have repulsive interactions. For this reason, these systems are modeled by soft-core bosons which are described most simply by the bose Hubbard model. Various aspects of this model were investigated analytically by mean-field theory \cite{Fishers,KampfZimanyi}, by renormalization group techniques \cite{Fishers,Rokhsar} and by projection methods \cite{Krauth0}. The bose Hubbard model has also been studied with quantum Monte Carlo methods (QMC) by Batrouni et al. \cite{Scalettar} in one dimension (1+1) and by Krauth and Trivedi \cite{Krauth3}, van Otterlo and Wagenblast\cite{vanOtterlo}, and Batrouni et al.\cite{Batrouni} in two dimensions (2+1). In this contribution, the Mott phase diagram is obtained from a strong-coupling expansion that has the correct dependence on spatial dimensionality, is as accurate as the QMC calculations, and agrees with the known exact solutions. Preliminary results for the pure case have already appeared\cite{Freericks_monien}. The bose Hubbard model is the minimal model which contains the key physics of the strongly interacting bose systems---the competition between kinetic and potential energy effects. Its Hamiltonian is \begin{equation} H = - \sum_{ij} t_{ij} b^\dagger_i b^{\phantom{\dagger}}_j+\sum_i\epsilon_i {\hat n}_i - \mu \sum_i {\hat n_i} + \frac{1}{2} U \sum_i {\hat n}_i ({\hat n}_i-1) \quad , \quad {\hat n}_i = b^\dagger_i b^{\phantom{\dagger}}_i \label{H} \end{equation} where $b_i$ is the boson annihilation operator at site $i$, $t_{ij}$ is the hopping matrix element between the site $i$ and $j$, $\epsilon_i$ is the local site energy, $U$ is the strength of the on-site repulsion, and $\mu$ is the chemical potential. The hopping matrix is assumed to be a real symmetric matrix $(t_{ij}=t_{ji})$ and the lattice is also assumed to be bipartite; {\it i.~e.}, the lattice may be separated into two sublattices (the $A$ sublattice and the $B$ sublattice) such that $t_{ij}$ vanishes whenever $i$ and $j$ both belong to the same sublattice (in particular, this implies $t_{ii}=0$). The local site energy $\epsilon_i$ is a quenched random variable chosen from a distribution of site energies that is symmetric about zero and satisfies $\sum_i\epsilon_i=0$. The pure case corresponds to all site energies vanishing $(\epsilon_i=0)$. The form of the zero temperature ($T=0$) phase diagram can be understood by starting from the strong-coupling or ``atomic'' limit \cite{Fishers,GiamarchiSchulz,Ma}. In this limit, the kinetic energy vanishes ($t_{ij} = 0$) and every site is occupied by a fixed number of bosons, $n_0$. In the pure case, the ground-state boson occupancy ($n_0$) is the same for each lattice site, and is chosen to minimize the on-site energy. If the chemical potential, $\mu=(n_0+\delta)U$, is parameterized in terms of the deviation, $\delta$, from integer filling $n_0$, then the on-site energy is $E(n_0) = -\delta U n_0 - \frac{1}{2} U n_0 (n_0 + 1)$, and the energy to add a boson onto a particular site satisfies $E(n_0+1) - E(n_0) = -\delta U$. Thus for a nonzero $\delta$, a finite amount of energy is required to move a particle through the lattice. The bosons are incompressible and localized, which produces a Mott insulator. For $\delta = 0$, the ground-state energies of the two different boson densities are degenerate [$E(n_0) = E(n_0+1)$] and no energy is needed to add or extract a particle; {\it i. e.}, the compressibility is finite and the system is a conductor. As the strength of the hopping matrix elements increases, the range of the chemical potential $\mu$ about which the system is incompressible decreases. The Mott-insulator phase will completely disappear at a critical value of the hopping matrix elements. Beyond this critical value of $t_{ij}$ the system is a superfluid. In the disordered case, a Mott-insulating phase may or may not exist depending upon the strength of the disorder. The energy to add a boson onto site $i$ becomes $E(n_0+1)-E(n_0)=\epsilon_i-\delta U$, so that the system is compressible if a site $i$ can be found which satisfies $\epsilon_i=\delta U$. If the disorder is assumed to be symmetrically bounded about zero ($|\epsilon_i|\le \Delta U$) then a Mott insulator exists whenever $\Delta <{1\over 2}$. The ground-state boson occupancy is uniformly equal to $n_0$ within the Mott insulating phase which extends from $-\Delta\ge \delta\ge\Delta-1$ (when $t_{ij}=0$). Once again, the bosons are incompressible within the Mott phase and the system is insulating. As the hopping matrix elements increase in magnitude, the range of the chemical potential within which the system is incompressible decreases until the Mott phase vanishes at a critical value of the hopping matrix elements. The compressible phase will typically also be an insulator and is called a bose glass\cite{Fishers}, but it has been conjectured that in some cases the transition proceeds directly from the Mott insulator to the superfluid\cite{Fishers,Rokhsar}. The phase boundary between the incompressible phase (Mott insulator) and the compressible phase (superfluid or bose glass) is determined here in a strong-coupling expansion by calculating both the energy of the Mott insulator and of a defect state (which contains an extra hole or particle) in a perturbative expansion of the single-particle terms $-\sum_{ij}t_{ij}b_i^{\dag}b_j+\sum_i\epsilon_i{\hat n}_i$. At the point where the energy of the Mott state is degenerate with the defect state, the system becomes compressible. In the pure case, the compressible phase is also superfluid, but in the disordered case, the compressible phase is a bose glass (except possibly at the tip of the Mott lobe)\cite{Fishers,Rokhsar}. There are two distinct cases for the defect state: $\delta < 0$ corresponds to adding a {\it particle} to the Mott-insulator phase (with $n_0$ bosons per site); and $\delta>0$ corresponds to adding a {\it hole} to the Mott-insulator phase (with $n_0+1$) bosons per site. Of course, the phase boundary depends upon the number of bosons per site, $n_0$, of the Mott insulator phase. To zeroth order in $t_{ij}/U$ the Mott-insulating state is given by \begin{equation} |\Psi_{\text{Mott}}(n_0)\rangle^{(0)} = \prod_{i=1}^N \frac{1}{\sqrt{n_0!}}\left(b^\dagger_i\right)^{n_0}|0\rangle \end{equation} where $n_0$ is the number of bosons on each site, $N$ is the number of sites in the lattice and $|0\rangle$ is the vacuum state. The defect state is characterized by one additional particle (hole) which moves coherently throughout the lattice. To zeroth order in the single-particle terms the wave function for the ``defect'' state is determined by degenerate perturbation theory: \begin{eqnarray} |\Psi_{\text{Def}}(n_0)\rangle^{(0)}_{\text{part}} &=& \frac{1}{\sqrt{n_0+1}} \sum_i f_i^{\rm (part)} b^\dagger_i |\Psi_{\text{Mott}}(n_0)\rangle^{(0)} \cr |\Psi_{\text{Def}}(n_0)\rangle^{(0)}_{\text{hole}} &=& \frac{1}{\sqrt{n_0}} \sum_i f_i^{\rm (hole)} b^{\phantom{\dagger}}_i |\Psi_{\text{Mott}}(n_0)\rangle^{(0)} \end{eqnarray} where the $f_i$ is the eigenvector of the corresponding single-particle matrix $S_{ij}^{\rm (part)}(n_0)\equiv -t_{ij}+\delta_{ij}\epsilon_i/(n_0+1)$ [$S_{ij}^{\rm (hole)}(n_0)\equiv -t_{ij}-\delta_{ij}\epsilon_i/n_0$] with the lowest eigenvalue (the hopping matrix is assumed to have a nondegenerate lowest eigenvalue). It is well known that the minimal eigenvalue of the single-particle matrix $S_{ij}$ is larger than the sum of the minimal eigenvalue of the hopping matrix plus the minimal eigenvalue of the disorder matrix. However, it has been demonstrated that as the system size becomes larger and larger, the minimal eigenvalue approaches the sum of the minimal eigenvalues of the hopping matrix and of the disorder matrix as closely as desired\cite{Lifshitz} (because of the existence of arbitrarily large ``rare regions'' where the system looks pure with $\epsilon_i=- \Delta U$ or with $\epsilon_i=\Delta U$). Therefore, in the thermodynamic limit, the perturbative energy of each defect state becomes \begin{equation} E_{\rm Def}^{\rm (part)}(n_0)-E_{\rm Mott}(n_0)=-\delta^{\rm (part)}U+ \lambda_{\rm min}(n_0+1)-\Delta U+... \label{eq: particle first order} \end{equation} \begin{equation} E_{\rm Def}^{\rm (hole)}(n_0)-E_{\rm Mott}(n_0)=\delta^{\rm (hole)}U+ \lambda_{\rm min}n_0-\Delta U+... \label{eq: hole first order} \end{equation} to first order in $S$, where $\lambda_{\rm min}$ is the minimal eigenvalue of the hopping matrix $-t_{ij}$. In the case of nearest-neighbor hopping on a hypercubic lattice in $d$-dimensions, the number of nearest neighbors satisfies $z=2d$ and the minimal eigenvalue is $\lambda_{\rm min}= -zt$. The boundary between the incompressible phase and the compressible phase is determined when the energy difference between the Mott insulator and the defect state vanishes (the compressibility is assumed to approach zero continuously at the phase boundary). Thus two branches of the Mott lobe can be found depending upon whether the defect state is an additional hole or an additional particle. The two branches of the Mott-phase boundary meet when \begin{equation} \delta^{\rm (part)}(n_0) + 1 = \delta^{\rm (hole)}(n_0). \label{eq: critical condition} \end{equation} The additional one on the left hand side arises because $\delta$ is measured from the point $\mu/U = n_0$. Equation (\ref{eq: critical condition}) may be used to estimate the critical value of the hopping matrix element beyond which no Mott-insulator phase exists. Let $x$ denote the combination $dt/U$ and consider the first-order expansions in Eqs. (\ref{eq: particle first order}) and (\ref{eq: hole first order}). The critical value of $x$ satisfies \begin{equation} x_{\rm crit}(n_0)={1-2\Delta\over 2(2n_0+1)}\quad , \label{eq: crit cond first order} \end{equation} which vanishes when the disorder strength becomes too large ($\Delta\ge 1/2$). Note that the critical value of $x$ is {\it independent} of the dimension of the lattice; {\it the dimensionality first enters at second order in $t$.} The slope of the phase boundaries about the point $\mu=n_0 U$ are equal in magnitude $[\lim_{x\rightarrow 0}\frac{d}{dx}\delta^{\rm (part)}(n_0,x) = - \lim_{x\rightarrow 0}\frac{d}{dx}\delta^{\rm (hole)}(n_0+1,x) ]$, but change their magnitude as a function of the density $n_0$, {\it implying that the Mott-phase lobes always have an asymmetrical shape}. Note further that the presence of disorder shifts the phase boundaries uniformly by $\Delta$, but the slope is {\it independent} of the disorder distribution. The bose Hubbard model in the absence of disorder is examined by a strong-coupling expansion through third order in the single-particle matrix $S$ in Section II. The exact solution for an infinite-dimensional lattice\cite{Fishers} is examined and various different extrapolation techniques are employed that do and do not utilize additional information of the scaling analysis of the critical point. Section III describes the similar results for the disordered bose Hubbard model and a discussion follows in Section IV. \section{The pure case} The bose Hubbard model in Eq. (\ref{H}) is studied in the absence of disorder ($\epsilon_i=0$). The many-body version of Rayleigh-Schr\" odinger perturbation theory is employed throughout. To third order in $t_{ij}/U$, the energy of the Mott state with $n_0$ bosons per site becomes \begin{equation} E_{\text{Mott}}(n_0) = N \left[ -\delta U n_0 - \frac{1}{2}U n_0 (n_0+1) - \frac{1}{N}\sum_{ij}\frac{t_{ij}^2} {U}n_0(n_0+1) \right] \label{eq: EMott} \end{equation} which is proportional to the number of lattice sites $N$. Note that the odd-order terms in $t_{ij}/U$ vanish in the above expansion (odd-order terms may enter for nonbipartite lattices). The energy difference between the Mott insulator and the defect state with an additional particle ($\delta < 0$) satisfies \begin{eqnarray} E_{\text{Def}}^{\rm (part)}(n_0) &- E_{\text{Mott}}(n_0) = -\delta^{(\text{part})} U+\lambda_{\rm min}(n_0+1) +{1\over 2U}\sum_{ij}t_{ij}^2f_j^2n_0(5n_0+4) -{1\over U}\lambda_{\rm min}^2n_0(n_0+1)\cr &+{1\over U^2} n_0(n_0+1) \Biggr[(2n_0+1)\lambda_{\rm min}^3-({25\over 4}n_0+{7\over 2}) \lambda_{\rm min}\sum_{ij}t_{ij}^2f_j^2-(4n_0+2)\sum_{ij} f_it_{ij}^3f_j\Biggr] \label{eq: Edef upper} \end{eqnarray} to third order in $t_{ij}/U$; while the energy difference between the Mott insulating phase and the defect phase with an additional hole ($\delta > 0$) satisfies \begin{eqnarray} E_{\text{Def}}^{\rm (hole)}(n_0) &- E_{\text{Mott}}(n_0)= \delta^{(\text{hole})}U+\lambda_{\rm min}n_0 + {1\over 2U}\sum_{ij}t_{ij}^2f_j^2(n_0+1)(5n_0+1) -{1\over U}\lambda_{\rm min}^2n_0(n_0+1)\cr &+{1\over U^2} n_0(n_0+1) \Biggr[(2n_0+1)\lambda_{\rm min}^3-({25\over 4}n_0+\frac{11}{4})\lambda_{\rm min}\sum_{ij}t_{ij}^2f_j^2-(4n_0+2)\sum_{ij}f_it_{ij}^3f_j\Biggr] \label{eq: Edef lower} \end{eqnarray} The eigenvector $f_i$ is the minimal eigenvector of the hopping matrix $-t_{ij}$ with eigenvalue $\lambda_{\rm min}$ and is identical in the particle and hole sectors. These results have been verified by small-cluster calculations on two and four-site clusters. Note that the energy difference in Eqs. (\ref{eq: Edef upper}) and (\ref{eq: Edef lower}) is {\em independent} of the lattice size $N$ indicating that QMC simulations should not have a very strong dependence on the lattice size. In the case of nearest-neighbor hopping on a $d$-dimensional hypercubic lattice, the minimum eigenvalue satisfies $\lambda_{\rm min}=-zt$, the sum $\sum_{ij}t_{ij}^2f_j^2$ becomes $zt^2$, and the sum $\sum_{ij}f_it_{ij}^3f_j$ equals $zt^3$. Equations (\ref{eq: Edef upper}) and (\ref{eq: Edef lower}) can then be solved for the shift in the chemical potential $\delta$ at which the system becomes compressible as a function of the parameter $x\equiv dt/U$. The results for the upper boundary are given by \begin{eqnarray} \delta^{(part)}(n_0,x) &= - 2x ( n_0 + 1) +{1\over d}x^2n_0(5n_0+4)-4x^2n_0(n_0+1)\cr &+2x^3n_0(n_0+1) \left [ (-8+{25\over 2d}-\frac{4}{d^2})n_0+(-4+{7\over d}-{2\over d^2}) \right ] \label{eq: upper boundary} \end{eqnarray} to third order in $x$, and the lower boundary is given by \begin{eqnarray} \delta^{(hole)}(n_0, x) &= 2xn_0-{1\over d}x^2(n_0+1)(5n_0+1)+4x^2n_0(n_0+1)\cr &-2x^3n_0(n_0+1) \left [ (-8+{25\over 2d}-{4\over d^2})n_0+(-4+{11\over 2d}-{2\over d^2}) \right ], \label{eq: lower boundary} \end{eqnarray} to third order in $x$. As a further check on the accuracy of the Mott phase boundaries in Eqs. (\ref{eq: upper boundary}) and (\ref{eq: lower boundary}), we compare the perturbative expansion to the exact solution on an infinite-dimensional hypercubic lattice\cite{Fishers} (which corresponds to the mean-field solution). Note that the solution in Ref.~\onlinecite{Fishers} was for the infinite-range-hopping model; this solution is {\it identical} to that on an infinite-dimensional lattice in the pure case. The Mott phase boundary may be expressed as \begin{equation} {\mu\over U}-n_0=-\frac{1}{2}-x\pm \sqrt{x^2-x(2n_0+1)+\frac{1}{4}} \label{eq: inf-d delta} \end{equation} where the plus sign denotes the upper branch to the Mott lobe $(\delta^{\rm (part)})$, and the minus sign corresponds to the lower branch $(\delta^{\rm (hole)}-1)$. The critical point can also be determined as the value of $x$ where the square root vanishes. One finds \begin{equation} x_{\rm crit}=n_0+\frac{1}{2}-\sqrt{n_0(n_0+1)} \label{eq: inf-d xcrit} \end{equation} which depends on $n_0$ as $1/n_0$ in the limit of large $n_0$. The strong-coupling expansions (\ref{eq: upper boundary}) and (\ref {eq: lower boundary}) agree with the exact solution (\ref{eq: inf-d delta}) when the latter is expanded out to third order in $x$, providing an independent check of the algebra. Note further, that the exact solution uniquely determines the expansion coefficients of the powers of $x$ that do not involve inverse powers of $d$ and the perturbation expansion is only required to determine the $1/d$ corrections. The strong-coupling expansion for the $x$, $\mu$ phase diagram in one dimension is compared to the QMC results of Batrouni et al. \cite{Scalettar} in Figure 1. The solid lines indicate the phase boundary between the Mott-insulator phase and the superfluid phase at zero temperature as calculated from Eq.~(\ref{eq: upper boundary}) and Eq.~(\ref{eq: lower boundary}). The solid circles are the results of the QMC calculation\cite{Scalettar} at a small but finite temperature ($T\approx U/2$). The dotted line is an extrapolation from the series calculation that will be described below. Note that the overall agreement of the two calculations is excellent. For example, the critical value of the hopping matrix element for the first Mott lobe ($n_0$) is $x_{\rm crit} = 0.215$, while the QMC calculations found\cite{Scalettar} $x_{\rm crit} = 0.215 \pm 0.01$. A closer examination of Fig. 1 shows a systematic deviation of the lower branch for larger values of $x$. We believe that this is most likely a finite-temperature effect, since the Mott-insulator phase becomes more stable at higher temperatures \cite{KampfZimanyi}, and the systematic errors of the QMC calculation due to finite lattice size and finite Trotter error are easily controlled\cite{Scalettar2}. It is known from the scaling theory of Fisher et al. \cite{Fishers} that the phase transition at the tip of the Mott lobe is in the universality class of the $(d+1)$ dimensional $XY$ model. Although a finite-order perturbation theory cannot describe the physics of the tricritical point correctly, we find that the density fluctuations dominate the physics of the phase transition even close to the tricritical point. Note how the Mott lobes have a cusp-like structure in one dimension, mimicking the Kosterlitz-Thouless behavior of the critical point. Figure 2 (a) presents the strong-coupling expansion for the $x$, $\mu$ phase diagram in two dimensions. For comparison, the tricritical point of the first Mott-insulator lobe as obtained by the QMC simulations of Krauth and Trivedi \cite{Krauth3} is marked by a solid circle with error bars (the chemical potential for the tip of the Mott lobe was not reported in Ref.~\onlinecite{Krauth3}, so we fixed it to be $\mu_{crit}$). The solid line is the strong-coupling expansion truncated to third order, while the dotted line is an extrapolation described below. Their simulation gives a critical value of $x_{\rm crit} = 0.122\pm 0.006$, whereas our calculation yields $x_{\rm crit} \approx 0.136$ which is in reasonable agreement. Note that the qualitative shape of the Mott lobes has changed from one dimension to two dimensions, mimicking the power-law critical behavior of the $XY$ model in three or larger dimensions. Figure 2 (b) shows the corresponding figure for the $n_0\rightarrow\infty$ limit corresponding to the quantum rotor model. The QMC results are from van Otterlo and Wagenblast\cite{vanOtterlo}. The horizontal axis has been rescaled to $y_{\infty}=\lim_{n_0\rightarrow\infty}n_0x$. We believe that the relatively large difference between the QMC and the strong-coupling perturbation theory arises from the use of the Villain approximation in the QMC simulations. Finally the strong-coupling expansion is compared to the exact calculation in infinite dimensions \cite{Fishers}. In infinite dimensions, the hopping matrix element must scale inversely with the dimension \cite{MuellerHartmann}, $t=t^*/d$, $t^* = \text{finite}$, producing the mean-field-theory result of Eq. (\ref{eq: inf-d delta}). In Figure 3 the strong-coupling expansion (solid line) is compared to the exact solution (dashed line) and to an extrapolated solution (dotted line) which will be described below. Even in infinite dimensions, the agreement of the strong-coupling expansion with the exact results is quite good. As a general rule, the truncated strong-coupling expansions appear to be more accurate in {\it lower} dimensions, which implies that the density fluctuations of the bose Hubbard model are also more important in lower dimensions. At this point we turn our attention to techniques which enable us to extrapolate the strong-coupling expansions to infinite order in hopes of determining a more accurate phase diagram. The simplest method is called critical-point extrapolation. The critical point $(\mu_{\rm crit},x_{\rm crit})$ is calculated at each order ($m$) of the strong-coupling expansion and is extrapolated to infinite order $(m\rightarrow \infty)$. The ansatz that the extrapolation is linear in $1/m$ can be checked by determining the correlation coefficient $r$ of the critical points (a value of $|r|$ that is near 1 indicates a linear correlation). The correlations are found to be most linear for large dimensions ($|r|=0.99999$ in infinite dimensions for the first Mott lobe) but remain fairly linear even in one dimension ($|r|>0.995$ for the $x_{\rm crit}$ extrapolation and $|r|>0.95$ for the $\mu_{\rm crit}$ extrapolation). Since the second- and third-order expansions are expected to be more accurate than the first-order calculation, we adopt the following strategy for performing the extrapolations: the results of the second- and third-order expansions are extrapolated to $m \rightarrow\infty$ to determine the estimate for the critical point, and the results of the first, second and third orders are then extrapolated to $m\rightarrow\infty$ in order to estimate the error in the critical point, and to test the linear-extrapolation hypothesis. The error estimate is chosen to be 1.5 times as large as the difference between the two different extrapolations. Figure 4 plots the critical hopping matrix elements $x_{\rm crit}$ versus $1/m$ for the infinite-dimensional case and $n_0=1,2,3$. The solid dots are the results of the strong-coupling expansion truncated to $m$th order and the solid line is the linear extrapolant. The open circles are the exact solutions from Eq. (\ref{eq: inf-d xcrit}). Note that although the linear correlation coefficient is very close to 1, the error in the critical point is about $2\%$. The results for the critical-point extrapolation are recorded in Table I. The critical-point extrapolation does not yield any information on the shape of the Mott lobes, but only determines the critical point. An alternate extrapolation technique, called the chemical-potential extrapolation method will determine an extrapolated Mott-phase lobe and critical point. The idea is to fix the magnitude of the hopping matrix elements and determine the value of the chemical potential from Eqs. (\ref{eq: upper boundary}) and (\ref{eq: lower boundary}) for the upper and lower branch of the Mott lobe. The chemical potential is determined from a first, second, and third-order calculation and then extrapolated to infinite order assuming the ansatz of a linear dependence upon $1/m$. This procedure determines an extrapolated Mott lobe that should be more accurate than the truncated strong-coupling series. The result for the infinite-dimensional case is presented as a dotted line in Fig. 3. Note that the critical point is not determined as accurately by this technique as it was in the critical-point extrapolation method. The chemical-potential extrapolation method fails in one dimension since the extrapolated branches of the extrapolated Mott lobe do not close. A third approach is to use the results of the scaling theory\cite{Fishers}. The critical point is that of a $(d+1)$-dimensional $XY$ model, and therefore, has a Kosterlitz-Thouless shape in one dimension and a power-law shape in higher dimensions. Examination of the exact result for infinite dimensions (\ref{eq: inf-d delta}) leads one to propose the following ansatz for the Mott lobe in $d\ge 2$ \begin{equation} {\mu\over U}-n_0=A(x)\pm B(x)(x_{\rm crit}-x)^{z\nu} \label{eq: mu ansatz} \end{equation} where $A(x)\equiv a+bx+cx^2+...$ and $B(x)\equiv \alpha +\beta x+\gamma x^2 + ...$ are regular functions of $x$ (that should be accurately approximated by their power-series expansions) and $z\nu$ is the critical exponent for the $(d+1)$-dimensional $XY$ model. In the unconstrained-scaling-analysis extrapolation method the exponent $z\nu$ is determined by the strong-coupling expansion in addition to the parameters $a,b,c$ and $\alpha,\beta,\gamma$. This provides a perturbative estimate of the exponent $z\nu$ which can be checked against its well-known values. In the constrained-scaling-analysis extrapolation method $z\nu$ is fixed at its predicted values\cite{Fishers} of $z\nu\approx 2/3$ in two dimensions and $z\nu=0.5$ in higher dimensions. In direct analogy to Eq. (\ref{eq: mu ansatz}), we propose the Kosterlitz-Thouless form \begin{equation} {\mu\over U}-n_0=A(x)\pm B(x)\exp \left [ -{W\over\sqrt{x_{\rm crit}-x}} \right ] \label{eq: mu ansatz 1d} \end{equation} for the constrained-extrapolation-method in one dimension. When the unconstrained-scaling-analysis extrapolation method is carried out, one finds that there is no solution for the critical exponent in one dimension (which is consistent with Kosterlitz-Thouless behavior), that in $d=2$ the exponent satisfies $z\nu\approx 0.58$, in $d=3$ the exponent is $z\nu\approx 0.54$, and in infinite dimensions $z\nu\approx 0.5$. There is a slight $n_0$ dependence to the exponents that are calculated in this method, but that arises from the truncation of the series to such a low order. In general, the unconstrained extrapolation method produces an accuracy of about $15\%$ in the exponent $z\nu$, and the method appears to work best in higher dimensions. The results for the constrained-extrapolation method are plotted with a dotted line in Fig. 1 for the one-dimensional case. The values of the critical points are $(\mu_{crit}=0.186, x_{crit}=0.265)$, $(1.319,0.155)$, and $(2.371,0.111)$ for $n_0=1,2,3$, respectively. These critical points occur at larger values of $x$ than predicted by the QMC simulations\cite{Scalettar}, but it is difficult to gauge whether the extrapolated series expansions are more or less accurate than the QMC simulations because of the finite-temperature effects in the latter. The constrained-extrapolation results in two dimensions are plotted with a dotted line in Fig. 2 (a) and (b). The values of the critical points are $(0.375,0.117)$, $(1.426,0.069)$, $(2.448,0.049)$ for $n_0=1,2,3$, respectively. The agreement with the QMC simulations\cite{Krauth3} is excellent. Similarly, the extrapolated critical point for the $2-d$ rotor model is $(0.5,0.171)$ which also agrees well with the QMC. In this latter case [Fig.~2~(b)] the errors between the extrapolated series expansion and the QMC can be traced to the use of the Villain approximation in the latter. When the constrained-scaling-analysis extrapolation method is applied to the infinite-dimensional case the result is indistinguishable from the exact solution when the two are plotted on the same graph. The extrapolation techniques work best in {\it higher} dimensions virtually producing the exact result in infinite-dimensions. This gives us confidence that the extrapolated results of the series expansions can produce numerical answers that are at least as accurate as the QMC simulations. \section{The disordered case} The most common type of disorder distribution that has been considered in relationship to the ``dirty'' boson problem is the Anderson model (continuous disorder distribution). In the Anderson model the distribution $\rho(\epsilon)$ for the on-site energies $\{\epsilon_i\}$ is continuous and flat, satisfying \begin{equation} \rho(\epsilon )=\theta(\Delta-\epsilon)\theta(\Delta+\epsilon) {1\over 2\Delta}\quad . \label{eq: anderson} \end{equation} The symbol $\Delta$ denotes the maximum absolute value that the site energy $\epsilon_i$ can assume for a given (bounded) distribution $(|\epsilon_i|\le \Delta U)$. This disorder distribution is symmetric $[\rho(-\epsilon)=\rho (\epsilon)]$ and in particular it satisfies $\sum_i\epsilon_i=0$. The results presented in this contribution are {\it insensitive} to the actual shape of the disorder distribution; all we require is a symmetric distribution with $|\epsilon_i|\le \Delta U$. We begin by reexamining the exact solution of the infinite-range-hopping model\cite{Fishers}. If all energies are measured in units of the boson-boson repulsion $U$, then the analysis of Ref.~\onlinecite{Fishers} derives an equation that relates the hopping matrix elements to the chemical potential at the Mott phase boundary \begin{equation} {1\over 2x} = \int_{-\infty}^{\infty} \left [ -{n_0+1\over y+\epsilon}+{n_0\over y+1+\epsilon} \right ] \rho(\epsilon ) d\epsilon\quad , \label{eq: inf-d disorder} \end{equation} with $y\equiv -n_0+\mu/U$ the chemical potential and $\rho (\epsilon )$ the disorder distribution. This solution assumes that the phase transition from the Mott phase to the bose glass is a {\it second-order} phase transition. When Eq. (\ref{eq: inf-d disorder}) is solved for the Anderson model distribution (\ref{eq: anderson}), one finds that the chemical potential for the lower branch of the Mott lobe behaves like $y=-\Delta-\exp[-1/2x(n_0+1)]$ for small $x$. This result is {\it nonperturbative} in the hopping matrix elements and cannot be represented by a simple perturbative theory about $x=0$. The reason why this happens is that the infinite-range-hopping model has no localized states for any disorder distribution (however, this statement does depend on the disorder distribution\cite{zimanyi}). Localized states can occur in the infinite-dimensional limit at the tails of the distribution. Therefore we expect that the transition will have a different qualitative character on a hypercubic lattice with nearest-neighbor interactions. In fact, the perturbative arguments given in the introduction, show that the phase boundaries have the {\it same} slope as they did in the pure case. Furthermore, we expect that the transition to be first order at the tip of the Mott lobe because the states that the bosons initially occupy in the compressible phase are localized within the rare regions of the lattice (where the site energies are all equal to $-\Delta U$) implying that there is no diverging length scale at the transition. The perturbative expansion for the energy of the Mott phase is unchanged from Eq. (\ref{eq: EMott}) in the presence of disorder (if the disorder distribution satisfies $\sum_i\epsilon_i=0)$, while the defect phases have a trivial dependence upon disorder (in the thermodynamic limit)---the energy for a particle or hole defect state is shifted by $-\Delta U$, so the effect of the disorder is simply to shift the Mott-phase boundaries inward by $\Delta U$. The critical point, where the Mott phase disappears, is no longer described by a second-order critical point [in which the slope of $\mu(x)$ becomes infinite at $x_{crit}$] but rather is described by a first-order critical point [in which the slope of $\mu(x)$ changes discontinuously at $x_{crit}$]. In the thermodynamic limit one can always find a rare region of arbitrarily large extent which guarantees the existence of the first-order transition, but the density of these rare regions is an exponentially small function of their size. For this reason the compressibility at the Mott-phase boundary will also be exponentially small. Finite-size effects play a much more important role in the disordered case: {\it it is impossible to determine the Mott-phase boundary accurately in the thermodynamic limit by scaling calculations performed on small lattices, because the lattice size must be large enough to contain rare regions within which the bosons can be delocalized.} (Finite-size effects can be studied with the strong-coupling expansion which is given to third order in the single-particle matrix $S$ in the appendix.) The most accurate way of calculating the Mott phase boundary is then to take the results of the constrained-scaling-analysis extrapolation for the pure case and shift the branches by the strength of the disorder. This is plotted in Figure 5 for the one-dimensional case and two different values of disorder ($\Delta=0,0.25$). The thermodynamic limit is represented by the solid line for the pure case, and dotted lines for the disordered case, while the dashed line is the result of an Anderson-model disorder distribution on a finite lattice with 256 sites. The QMC results of Batrouni, {\it et.~al.}\cite{Scalettar} correspond to lattice sizes ranging from 16 sites to 256 sites (the disorder parameter is $\Delta=0$ for the solid dots and $\Delta=0.25$ for the open dots). The Mott phase is stabilized on finite-sized systems because they do not possess the rare regions needed to correctly determine the Mott phase boundary. This is clearly seen in the QMC results, which predict a much larger region for the Mott phase than the strong-coupling perturbation theory does in the thermodynamic limit. The perturbative results for a finite system are much closer to the QMC results as expected. (Note that the finite-size calculations have not been extrapolated, so they should underestimate the stability of the Mott phase in one dimension which is clearly seen in Figure~5.) Also the slope of the phase boundary approaches zero (as $x\rightarrow 0)$ for the finite-size systems\cite{Scalettar}. Figure 6 plots the Mott-phase diagram for the disordered bose Hubbard model in two dimensions and one value of the disorder ($\Delta=0,0.182$) in the thermodynamic limit. The solid dot (with error bars) is the QMC result\cite{Krauth3} (for the pure case with $\Delta=0$) and the open dot is the disordered case $(\Delta=0.182$). Note that in dimensions larger than one, the finite-size effects for the tip of the Mott lobe are not as strong as they are in one dimension. We compare in Figure 7 the differences between the infinite-dimensional lattice and the infinite-range-hopping model of Ref.~\onlinecite{Fishers}. The solid lines correspond to the exact solution with no disorder, the dotted lines are the infinite-dimensional lattice strong-coupling expansion with disorder ($\Delta=0.2$), and the dashed lines are the exact solution of the infinite-range-hopping model. In the case of disorder, the first-order nature of the transition is evidenced by the jump in the slope of the Mott phase boundary at the tip of the lobe. The second-order phase boundaries predict a more stable Mott phase, and their slopes all approach zero as $x\rightarrow 0$. We expect in the region in between the (first-order) infinite-dimensional phase boundary and that of the (second-order) infinite-range-hopping model that the compressibility will be exponentially small, and will only become sizable as the second-order phase boundary is crossed. Because the Mott-phase to bose-glass phase transition is first order for the disordered case, and since the bose-glass to superfluid transition is always second order (because it involves a collective excitation that extends through the entire lattice), it is quite unlikely that there would ever be a region where the Mott phase has a transition directly to the superfluid. {\it The presence of the Lifshitz rare regions strongly supports the picture that the Mott phase is entirely enclosed within the bose-glass phase.} This result is {\it independent} of any perturbative arguments, since the rare regions must dominate the Mott to bose-glass transition in the exact solution too. Finally, we calculate the dependence of the critical value of $x$ at the tip of the Mott-phase lobe, as a function of the disorder strength $\Delta$. Figure 8 plots this value of $x_{crit}(\Delta)/x_{crit}(0)$ versus $\Delta$ for the one-, two-, and infinite-dimensional cases. The plot is limited to the lowest Mott-phase lobe with $n_0=1$. Since the one-dimensional Mott phase lobes have a cusp-like behavior that is removed when disorder is added to the system, we expect $x_{crit}$ to decrease very rapidly for small disorder. This effect is sharply reduced in higher dimensions. In the strong-disorder limit, the phase diagram is dominated by the first-order terms in the perturbative expansion, which have a trivial dependence on the dimensionality, but the slopes curves of the curves are unequal because $x_{\rm crit}(0)$ depends strongly upon the dimensionality. \section{Conclusion} We have developed a strong-coupling perturbation-theory approximation to the bose Hubbard model on a bipartite lattice. The perturbative results can be extrapolated in a number of different ways which either do or do not take into account the scaling theory of the critical point at the Mott tip. We find that a perturbative expansion thru third order rivals the accuracy of the QMC simulations, and is likely the best method for determining the Mott phase boundary of these interacting bose systems. We treated two different cases: the pure case and the disordered case. In the pure case the tip of the Mott lobe satisfies a scaling relation because it corresponds to a second-order phase transition in a $d+1$-dimensional XY model. This is because the compressible phase is also superfluid which implies there is a diverging length scale at the phase transition. Calculations in the pure case are insensitive to finite-size effects. In the disordered case we argued that the tip of the Mott phase lobe corresopnds to a first-order phase transition because the initial single-particle excitations are localized into the rare regions of the Lifshitz tails for any bounded disorder distribution. As a result there is a kink in the Mott phase boundary since the slope of $\mu(x)$ has a discontinuous jump at the tip of the lobe. In this case, there are strong finite-size effects because ``typical'' disorder distributions on finite lattices do not have Lifshitz tails. The results of these perturbative calculations have been compared to the available QMC simulations. We find a remarkable agreement between the two and are unable to determine which method is more accurate in a quantitative determination of the phase boundaries. The perturbation theory described here falls short in one aspect---it is unable to determine the bose-glass--superfluid phase transition in the disordered case. It is possible that such a calculation could be performed, but since the particle density at which it occurs is a priori not known, and since the superfluid susceptibility diverges in the bose-glass phase, such a calculation may be problematic. \acknowledgments We would like to thank M. Fisher, Th. Giamarchi, M. Jarrell, M. Ma, A van Otterlo, R. Scalettar, R. Singh, and G. Zimanyi for many useful discussions. We would especially like to thank M. Ma for pointing out that the Mott to bose-glass phase transition is first order in the disordered case. Initial stages of this work were carried out by J. K. F. at the University of California, Davis in 1994 and were completed during a visit to l'\'Ecole Polytechnique F\'ed\'erale de Lausanne in June 1995. JKF would like to thank the Office of Naval Research (under Grant No. N00014-93-1-0495) for support while at UC Davis, and would like to thank the Donors of the Petroleum Research Fund, administered by the American Chemical Society (ACS-PRF-29623-GB6) for support while at Georgetown.
\section{acknowledgments} AG acknowledges support from the HCM program under contract ERB4001GT932058. This work was supported by NATO, ONR (USA), INFN (Italy), EPSRC (UK), The Fulbright Foundation and The Donors of the Petroleum Research Fund administered by The American Chemical Society.
\subsection{The scalar Bethe-Salpeter equation} The BS equation for bound state of two equal mass scalar fields, "quarks", interacting via scalar exchanges has a form: {\begin{eqnarray} \psi (p,P) = D(p_1,p_2)^{-1}\int \frac{d^4 p'}{(2\pi)^4}{\cal K}(p,p',P)\psi(p',P), \label{sbse} \end{eqnarray} } where $\psi(p,P)$ is the BS amplitude for the bound state with momentum $P$, ${\cal K}(p,p',P)$ is a kernel of BS equation and $D(p_1,p_2)^{-1}$ is the two-body propagator. Generally speaking, the kernel ${\cal K}$ is a sum of the all renormalized irreducible graphs presenting the Green's function with two incoming and two outcoming constituent fields and the propagator $D^{-1}$ is product of the renormalized full single-particle propagators. In practice, however, both of them are usually taken in the lowest order in the coupling constant. It corresponds to the ladder approximation for the kernel with free particle propagators. In the simplest case of an exchange field of mass $\mu$ and coupling constant $g$, the ladder approximation for the kernel has an explicit form: {\begin{eqnarray} {\cal K}_{ladder}(\mu,p,p',P) = {\cal K}_{ladder}(\mu,p,p') \equiv \frac{{\it i}g^2}{(p-p')^2-\mu^2}, \label{ladk} \end{eqnarray} } The two-body propagator then reads {\begin{eqnarray} D(p_1,p_2)\equiv d(p_1)\cdot d(p_2) = \left ({p_1^2-m^2}\right )\cdot \left ({p_2^2-m^2}\right ), \label{twoprop} \end{eqnarray} } where $d(p)^{-1}$ is free one-particle propagator, $m$ is the mass of the constituents and $p_{1,2} = \frac{1}{2}P \pm p$ are the constituent momenta. Phenomenological applications of the equation (\ref{sbse}) (or more interesting spinor-spinor equation) requires a more general form of the kernel than eq. (\ref{ladk}). More degrees of freedom are provided by introducing a sum of the exchanges with different parameters and different Lorents structure, e.g. vector or pseudoscalar exchanges. In the presence of the several (effective) exchanges with different quantum numbers and in higher orders kernel may contains both attractive and repulsive terms. Here we consider the kernel of the form: {\begin{eqnarray} {\cal K}_G(p,p') = \sum_{j=1}^{N} \epsilon_j \frac{{\it i} g_j^2}{(p-p')^2-\mu_j^2}, \label{kerg} \end{eqnarray} } where index $j$ enumerates different "exchange" terms, distinguished by the mass parameter, $\mu_j$, and strength, $g_j$. Factor $\epsilon_j=\pm 1$ defines attractive ($\epsilon_i=+1$) or repulsive ($\epsilon_j=-1$) type of the correspondent term. Being interpreted as contribution of the lowest order diagram, the terms with negative value of $\epsilon_j$ require either antihermitian term in the interaction hamiltonian (lagrangian) or fields with additional quantum numbers. However, we consider the entire sum in the eq. (\ref{kerg}) as a convenient parametrization of the unknown full kernel of the BS equation. The convenience of such a form of the kernel is that kernel is explicitly covariant and contains only "field theoretical" elements, the free particle propagators. Accepting this way of action, we assume that parameters of the kernel should be fixed to describe the experimentally known spectrum of the system. The free propagator $D(p_1,p_2)^{-1}$ of the form (\ref{twoprop}) and the kernel in the ladder approximation, (\ref{ladk}) or (\ref{kerg}), usually provide the basis for an investigation of the BS equation. The parameters for a phenomenological adjustment of the theoretical framework are the coupling constants, $g_j$, and exchange masses, $\mu_j$. The parameter $m$ in the propagator is referred to as "physical mass" of the constituent and it is supposed to include effectively the self-energy corrections. In the case of the observable particles, e.g. nucleons, $m$ is the observable mass, otherwise, e.g. quarks, it has rather ambiguous model-dependent value. In order to study the spectrum of the BS equation we, first, fix the frame of calculation as the rest frame of the system, where $P=(M,{\bf 0})$ and $M$ is the mass of the system, "meson". Then, we perform the well-known Wick rotation~\cite{wick}, which may later cause difficulties in the calculation of observables of some reactions, but does not affect the spectrum of the equation and simplifies the numerical analysis. Under the Wick rotation the BS equation keeps the form (\ref{sbse}) with the ladder kernel {\begin{eqnarray} {\cal K}_{ladder}(\mu,p,p') = \frac{g^2}{(p-p')^2_E+\mu^2},\quad (p-p')_E^2 = (p_0-p'_0)^2 +({\bf p}-{\bf p'})^2, \label{ladkr} \end{eqnarray} } and propagator {\begin{eqnarray} D(p,M) = {\left [ (p_0^2+m^2 + {\bf p}^2- M^2/4)^2+M^2 p_0^2 \right ]}. \label{twopropr} \end{eqnarray}} We use the previous notation for the "rotated" functions $\psi$, ${\cal K}$ and $D$, and for the "Euclidean" momenta, $p$ and $p'$. Next, we perform a partial wave decomposition of the equation~\cite{scale}: {\begin{eqnarray} \frac{1}{(p-p')_E^2+\mu^2}&=&\frac{2\pi}{\mid {\bf p}\mid\cdot\mid{\bf p'}\mid} \sum_{L=0}^{\infty} \sum_{\lambda=-L}^{L} Q_L(\beta) Y_{L\lambda}(\theta_p,\varphi_p) Y_{L\lambda}^*(\theta'_p,\varphi'_p),\label{part1}\\ \psi(p,M) &=&\frac{1}{\mid {\bf p}\mid} \sum_{L=0}^{\infty} \sum_{\lambda=-L}^{L} \psi_L(p_0,\mid {\bf p}\mid,M) Y_{L\lambda}(\theta_p,\varphi_p), \label{part} \end{eqnarray} } where we have already taken into account that $\psi_L$ are independent on the projection of angular momentum, $\lambda$; the dimensionless parameter $\beta$ defined by the expression $$ \beta = \frac{\mu^2 +{\bf p^2 + p'^2} + (p_0 -p'_0)^2} {2\mid {\bf p}\mid\cdot\mid{\bf p'}\mid}, $$ and $Q_l$ are the Legendre functions of the second kind. For $l=0,1$ they are: \begin{eqnarray} Q_0(y) = \frac{1}{2}{\sf ln}\left ( \frac{y+1}{y-1} \right ) , \quad\quad Q_1(y) = \frac{y}{2}{\sf ln}\left ( \frac{y+1}{y-1} \right ) - 1 .\label{legql} \end{eqnarray} Substituting (\ref{part1}) and (\ref{part}) into eq. (\ref{sbse}) and performing the angular integration, we arrive at a set of the independent equations for the partial amplitudes $\psi_L$, corresponding to the states with the angular momentum $L$: {\begin{eqnarray} &&\psi_L (p_0,|{\bf p}|,M) = D(p,M)^{-1}\int \frac{dp'_0 d|{\bf p'}|}{(2\pi)^3} {\cal K}^L(p_0,|{\bf p}|,p'_0,|{\bf p'}|) \psi_L(p'_0,|{\bf p'}|,M), \label{sbsem}\\[2mm] &&\quad\quad\quad\quad\quad\quad\quad\quad {\cal K}^L(p_0,|{\bf p}|,p'_0,|{\bf p'}|) = g^2 Q_L(\beta). \label{park} \end{eqnarray}} The amplitude of pure quantum state with defined angular momentum $L$ and its projection $\lambda$ is then given by {\begin{eqnarray} \psi(p,M,L,\lambda) = \psi_L( p'_0,|{\bf p'}|,M)Y_{L\lambda}(\theta_p,\phi_p). \label{pure} \end{eqnarray}} The two-dimensional integral equations (\ref{sbsem}) are an exact projection of the initial equation (\ref{sbse}) with (\ref{ladk}) and (\ref{twoprop}). In the case of the generalized kernel (\ref{kerg}), there will be a sum over different terms on the right hand side of eq. (\ref{sbsem}): {\begin{eqnarray} {\cal K}_{G}^L(p_0,|{\bf p}|,p'_0,|{\bf p'}|) = \sum_{j=1}^N \epsilon_j g_j^2 Q_L(\beta_j). \label{parks} \end{eqnarray} } To obtain the spectrum of the BS equation we need to solve the eigenvalue problem for the bound state mass $M$ at fixed set of parameters of the kernel. =It is to be noted that the BS equation is not linear in the mass $M$. On the other hand it is linear in the coupling constants squared, $g^2$. Therefore we consider an equation of the form: {\begin{eqnarray} \psi_L (p_0,|{\bf p}|,M) = \lambda \cdot D(p,M)^{-1}\int \frac{dp'_0 d|{\bf p'}|}{(2\pi)^3} {\cal K}_G^L (p_0,|{\bf p}|,p'_0,|{\bf p'}|)\psi_L(p'_0,|{\bf p'}|,M), \label{sbsel1} \end{eqnarray} } or symbolically {\begin{eqnarray} \psi = \lambda \cdot \hat {\cal K} \psi, \label{sbsel} \end{eqnarray} } where we skip indicies $L$ for shortness. Solving the linear eigenvalue problem for $\lambda$ at fixed parameters of the kernel and various values of the bound state mass, we can map $\lambda(M)$. Then by inverting the mapping as $M(\lambda)$, we will find solutions of the eigenvalue problem for $M$ with the kernel $\lambda\cdot \hat{\cal K}$, where factor $\lambda$ is trivially absorbed by redefining the coupling constants. Traditional methods to find solutions to the linear eigenvalue problem corresponding to the integral equation (\ref{sbsel}) are based on the ideas of the classical Fredholm theory. The basic idea here is to substitute integration by a summation and then deal with a sufficiently large system of linear algebraic equations. The problem is reduced to finding the eigenvalues of the matrix corresponding to the integral operator $\hat {\cal K}$. However, in the case of the covariant BS equation such a program of action meets extra difficulties in view of the double integration, which leads necessarily to very large matrices, say $\sim 10^4 \times 10^4$ or even larger in the case of spinor-spinor equation (with additional factor of $64$ if the parity is conserved by interactions and factor 256 if it does not). We use an alternate method based on the iteration of the BS equation in the form (\ref{sbsem}). This method in the essence is similar to the Malfliet-Tjon method~\cite{maltjon} employed to solve the integral Schr\"odinger equation in the nuclear physics (see also discussion in ref.~\cite{stadler}). \subsection{The iteration method} The iteration of any trial function, $\Phi^{(0)}$, with eq. (\ref{sbsem}), is understood as obtaining other function $\Phi^{(n)}$ using the algorithm (we use the symbolic notations as in eq. (\ref{sbsel})): {\begin{eqnarray} \Phi^{(i+1)} = \lambda \cdot \hat {\cal K} \Phi^{(i)}, \label{it1}\\ \Phi^{(n)} = \left [ \lambda \cdot \hat {\cal K} \right ]^n\Phi^{(0)}, \label{it2} \end{eqnarray} } In order to organize the iteration (\ref{it1})-(\ref{it2}) on the computer, we need an "integrator", corresponding to the operator $\hat {\cal K}$. This means that we have to perform the computer calculation of the double integral on the r.h.s. of the equation with a defined kernel and any trial function, $\Phi^{(0)}$, (we assume good enough behavior of the function, such as absence of singularities and vanishing at large arguments, $|{\bf p}| \to \infty$ and $p_0 \to \pm \infty$). In our particular calculations this integrator is organized as a two dimensional Gauss integration with suitable mapping of variables. Next, we would like to know what happens with equation after sufficiently large number, $n$, of iterations. Let us assume that solutions, $\psi_\alpha$, of the equation (\ref{sbsel}) corresponding to the eigenvalues $\lambda_\alpha$ belong to the complete system of functions\cite{foot2}. Therefore, the trial function can be expanded as {\begin{eqnarray} \Phi^{(0)} = \sum_{\alpha=0}^\infty A_\alpha \psi_\alpha. \label{it3} \end{eqnarray}} Thus, the result of iteration (\ref{it1})-(\ref{it2}) is: {\begin{eqnarray} \Phi^{(n)} = \sum_{\alpha=0}^\infty A_\alpha \left [\frac{ \lambda}{\lambda_\alpha} \right ]^n\psi_{\alpha}. \label{it4} \end{eqnarray}} {}From last equation it is obvious that at sufficiently large $n = N$ all terms with $\alpha > 0$ are small compared to the ground state term, $\alpha = 0$: {\begin{eqnarray} \lim\limits_{{n\to N}}\Phi^{(n)} \equiv \Phi^{(N)} = C \cdot \psi_{0} + {\cal O}\left (\left [ \frac{ \lambda_0}{\lambda_1} \right ]^N \right ) . \label{it5} \end{eqnarray}} Therefore $N$ to be chosen to make the last term on the r.h.s. of (\ref{it5}) to be negligibly small. Then, comparing $\Phi^{(N)}$ and $\Phi^{(N+1)}$, we find the ground state eigenvalue, $\lambda_0$: {\begin{eqnarray} \lambda_0 = \frac{\Phi^{(N)} }{ \Phi^{(N+1)}}. \label{it6} \end{eqnarray} } This recipe to find the ground state eigenfunction and eigenvalue works nicely numerically and, of course, does not depend on the choice of the initial trial function. Formula (\ref{it3}) also provides us with the possibility to find higher levels on $\lambda$. Indeed, taking the combination of iterating functions: {\begin{eqnarray} \Phi^{(i+1)} - \left [ \frac{ \lambda}{\lambda_0} \right ] \cdot \Phi^{(i)} = \sum_{\alpha=1}^\infty \left ( 1 - \frac{\lambda_\alpha}{\lambda_0}\right )A_\alpha \left [\frac{ \lambda}{\lambda_\alpha} \right ]^{(i+1)}\psi_{\alpha}, \label{it7} \end{eqnarray}} i.e. new trial function which does not contain admixture of the ground state! It is easy to see that iterations of this function give the eigenfunction $\psi_1$ of the first excited state, $\lambda_1$, similarly to the procedure for ground state. The same procedure, in principle, can be organized for any desired level on $\lambda$. Thus, the problem is solved. The only limitation is, of course, accuracy of the numeric calculations. Calculation of the high $\lambda$ require a precise calculation of all levels below, which leads to substantial computer time. No special numerical problems were find in calculating the three lowest levels for eq.~(\ref{sbsel}). The numerical results are presented in the next section. \subsection{Numerical results for iteration method} To study numerically the capability of the method to solve the integral BS equation we consider eq.~(\ref{sbsem}) with the model kernel of the form (\ref{parks}). The parameters of the model are presented in Table~I. We refer to the constituent fields as "quarks" and their mass, $m$, is chosen to be similar to the mass of the $c$-quark. The parameters of the kernel are chosen to provide the typical density of the levels of the $c\bar c$-bound state, the charmonium, not too far from the limit of the spectrum $M_{lim} = 2m$. \begin{center} \begin{tabular}{|c|c|c|} \hline coupling constants & $\epsilon_j $ & mass \\ $g^2_j/(4\pi), \quad GeV^2$ & &$\mu_j$, GeV \\ \hline \hline 37800.0 & +1 & 0.10 \\ \hline 37800.0 & -1 & 0.11 \\ \hline 45.0 & -1 & 0.95 \\ \hline 45.0 & +1 & 1.425 \\ \hline \hline \multicolumn{3}{|c|}{$m = 1.5 $ GeV, }\\ \hline \end{tabular} \vspace*{5mm} Table. I. The parameters of model ("scalar charmonium"). \end{center} Results of a calculation of the spectrum, $\lambda(M)$, are presented in Fig.~1 for the three lowest levels in the channels with $L=0$ (S-states) and $L=1$ (P-states). The physical spectrum $M(\lambda)$ can be obtained crossing the plot by the line $\lambda = \lambda_{phys}$. For instance, as is shown in Fig.~1 at $\lambda_{phys} = 0.13$ the masses of the lowest states are (with accuracy $\sim 0.5\%$): \begin{eqnarray} M(1S) = 1.265 \quad {\rm GeV}; \quad M(2S) =1.939 \quad {\rm GeV};\quad M(3S) = 2.251 \quad {\rm GeV}; \label{m1} \\ M(1P) =1.751 \quad {\rm GeV} ; \quad M(2P) =2.146 \quad {\rm GeV};\quad M(3P) = 2.385 \quad {\rm GeV}. \label{m2} \end{eqnarray} For illustartion, we pesent the amplitudes $\psi_0$, corresponding the spectrum (\ref{m1}) in Fig.~2. These amplitudes are shown as a functions of the spatial momentum $|{\bf p} |$ and at $p_0 = 0$. We see that the type of radial excitations is similar to the one for the Schr\"odinger equation. However, in general case of the BS equation we deal with the hyperradial excitations of two-dimensional surface, $\psi_L (p_0,|{\bf p} |)$. \section {Confinement for the Bethe-Salpeter equation} \subsection{General discussion and non-relativistic confining potentials} The idea of confinement has different realization within different theoretical approaches. The simplest intuitive picture is given by the non-relativistic bound state formalism based on the Schr\"odinger equation with a QCD inspared phenomenological potential. A system of two particles interacting in a non-confining potential, vanishing as $r \to \infty$, has a spectrum of bound states with an upper limit, $M_{lim} = 2m_q$, and a continuum above this limit. The confinement is conventionally associated with the infinitely rising linear part of the full $q\bar q$-potential, $V_l = \alpha r, \alpha > 0$, which provides with the mass spectrum extending infinitely beyond $M_{lim}$. It is clear that this mechanism can not be directly adopted by the covariant field theoretical approaches, such as the BS formalism. More relevant approach is based on the simultaneous analysis of the BS, Schwin\-ger-Dyson (SD) equations and Ward-Takahashi at the lowest order and with the model gluon propagator~\cite{munczek,williams}. In particular, the role of the analytical structure (structure of singularities) of the quark propagators is discussed here. It is found that, with certain choices of the model gluon propagator, the quark propagator is an entire function (function with no singularities) at physical momentum of the quark. This important property of the quark propagator is considered as an indication of confinement~\cite{qpr,williams,tw}. These two examples present two essentially different pictures of what is referred to as confinement. In the first case, the confinement is the two-body effect, i.e. quarks can not escape from each other because of the interaction between them. In the second example, confinement is attributed to the property of a single quark, which can not propagate as a free particle. The general formalism of the field theory suggests that the two-body and the single particle phenomena are in a generic relationship. So do the approaches based on the BS and SD equations. Our approach to a modeling of the confinement is in some sense inverse to that of refs.~\cite{qpr,williams}. We, first, construct the covariant kernel of the BS equation which, we expect, would provide the confinement and only then study the modifications of the quark propagators involved in the BS equation. We start from a few unsophisticated observations prompted by the non-relativistic picture: \begin{enumerate} \item In the non-relativistic limit the covariant theory can be reduced to the formalism with the Schr\"odinger equation, where we can expect the picture of the confinement as interaction with non-vanishing potential at $r\to \infty$ to be valid. The main distinguishing feature of the spectrum here is the existence of the bound states above the two quark mass limit, $M_{lim}$. \item It is not necessary to have an infinitely rising potential, if we intend to discuss only a few lowest levels in the spectrum. More manageable potential, $V \to V_{\infty} >0$ at $r\to \infty$, could be sufficient, if $V_{\infty}$ is large enough. In this sense, we also refer to the constant potential as a confining one. \item The potential in the momentum space $V({\bf k})$ can be obtained as a non-relativistic limit of the kernel ${\cal K}$, similar to \begin{eqnarray} V({\bf k}) = -\frac{g^2}{{\bf k}^2+ \mu^2} \label{yuk} \end{eqnarray} if the kernel is of the form (\ref{ladk}). Important point here is that the non-relativistic form of the potential (\ref{yuk}) is of a field-theoretic origin. \end{enumerate} Basing on the last observation we expect that if we define a way to construct a confining potential from the non-relativistic field theory, then we can apply similar methods to obtain the relativistic confining kernel. Very often the following recipe is used. The non-relativistic Yukawa potential in the coordinate space $V(r)$, the Fourier transform of eq. (\ref{yuk}), is \begin{eqnarray} V({\bf r}) = -\frac{g^2}{4\pi}\frac{e^{-\mu r}}{r}. \label{yukr} \end{eqnarray} The linear potential can be derived as \begin{eqnarray} V_l = \lim\limits_{\mu \to 0} \left [ -\frac{\partial^2}{\partial \mu^2}V({\bf r}) \right ]= \lim\limits_{\mu \to 0}\frac{g^2}{4\pi}{r}{e^{-\mu r}}=\frac{g^2}{4\pi}{r}. \label{dyukr} \end{eqnarray} The the relativistic generalization is made by a Fourier transform to the momentum space and replacing the non-relativistic Yukawa potential, (\ref{yuk}), by it relativistic analog: \begin{eqnarray} {\cal K} \propto \lim\limits_{\mu \to 0} \left [ -\frac{\partial^2}{\partial \mu^2} \frac{g^2}{ {k}^2- \mu^2} \right ]. \label{dyukp} \end{eqnarray} Taking the limit in (\ref{dyukp}), one should exercise great deal of care, since this leads to the appearance of generalized functions in the kernel~\cite{grossmilana,vary}. This recipe give us a guideline, however it is not completely satisfactory, since (i) the kernel (\ref{dyukp}) (or the potential (\ref{dyukr})) is not of the field-theoretic form and (ii) it is not clear does the direct use of operation (\ref{dyukp}) lead to the rising or, at least, non-vanishing interaction in the four-dimensional space. Intending to stay with our parametrization of the kernel as a superposition of the ladder terms, similar to (\ref{kerg}), we have to find an appropriate presentation of the operation (\ref{dyukr}). Let us start with a superposition of the non-relativistic potentials: \begin{eqnarray} V({\bf r}) = \sum_j \frac{C_j}{r}{\sf exp}[-\mu_j r] = \sum_j \frac{C_j}{r}{\sf exp}[-\mu \alpha_j r], \label{syukr} \end{eqnarray} where $\mu$ provides the mass scale and $\alpha_j$ are dimensionless parameters. Then expanding the exponents we get \begin{eqnarray} V({\bf r}) &=&\sum_j {C_j}\left [ \frac{1}{r} -\mu \alpha _j + \mu^2\frac{\alpha_j^2 r}{2} -\ldots \right ]. \label{syukr1} \end{eqnarray} {}From eq. (\ref{syukr1}) we see that, taking the limit $\mu \to 0$ and correspondly adjusting the parameters $C_j$ and $\alpha_j$, desired non-vanishing behavior of the potential can be provided. For instance, for constant potential, $V_c$, we have \begin{eqnarray} C_j = \mu^{-1}\tilde C_j; \quad \sum_j \tilde C_j = 0; \quad \sum_j \alpha_j\tilde C_j \equiv A_c\ne 0, \label{p0} \end{eqnarray} where we need only two terms to satisfy the conditions, i.e. $j_{max}=2$. For linear potential,$V_c$, \begin{eqnarray} C_j = \mu^{-2}\tilde C_j; \quad \sum_j \tilde C_j = 0; \quad \sum_j \alpha_j\tilde C_j = 0; \quad \sum_j \alpha_j^2\tilde C_j \equiv A_l\ne 0, \label{p1} \end{eqnarray} where $j_{max}=3$. The limit $\mu \to 0$ corresponds to the physical picture of the superposition of very light mass exchanges with slightly different masses and large coupling constants. Please, note that power of the non-vanishing term in the limit $\mu \to 0$ is solely controlled by the power of the $\mu$ in the denominator of the coefficients $C_j$. Using a superposition of the Yukawa potentials in the momentum space, we find (see Appendix~A): \begin{eqnarray} V_c ({\bf k}) &=& A_c \delta^{(3)}({\bf k}), \label{p0p}\\[3mm] V_l ({\bf k}) &=& \left( \frac{1}{2}-\frac{{\sf ln}2}{\pi} \right ) A_l\delta^{(3)}({\bf k})\frac{\partial }{\partial k}. \label{p1p} \end{eqnarray} Note that, for the linear potential, the Schr\"odinger equation becomes an integro-differential equation in momentum space. \subsection{Confining kernel for the Bethe-Salpeter equation}. We look for a confining kernel, ${\cal K}_{con}$, of the BS equation in the form of the superposition of ladder kernels in the momentum space and after the Wick rotation: \begin{eqnarray} {\cal K}_{con }(k_E) &=& \sum_j \frac{C_j}{k_E^2 + \alpha_j^2 \mu^2}, \label{conk} \end{eqnarray} where $ k_E^2 = k_0^2+{\bf k}^2$. In spite of the obvious similarity to the non-relativistic case, the expression for the relativistic kernel, (\ref{conk}), has essentially different properties in view of the larger dimension of the space. Indeed, the ladder kernel in the coordinate space is~\cite{kercoor}: \begin{eqnarray} {\cal K}_{ladder} (R_E,\mu) = g^2 \mu \frac{{\sf K}_1(\mu R_E) }{R_E}, \label{kerco} \end{eqnarray} where $R_E = ({\bf r}^2+t^2)^{1/2}$. The asymptote at small $R_E$ can be obtained as: \begin{eqnarray} \!\!\!\!\! \!\!\!\!\!\!\!\!\!\! \!\!\!\!\! && {\cal K}_{ladder} (R_E\to 0,\mu) \sim g^2 \left \{ \frac{1}{R_E^2} \right . \nonumber \\ \!\!\!\!\! \!\!\!\!\!\!\!\!\!\! \!\!\!\!\!&& \quad \left .+ \frac{\mu^2}{2} \sum_{m=0}^{\infty} \frac{1}{m!(m+1)!} \left ( \frac{\mu R_E}{2} \right )^m \left [ {\sf ln} \left (\frac{\mu R_E}{2} \right ) -\frac{1}{2} \left (\psi (m+1)+\psi(m+2)\right) \right] \right \}, \label{kerco1} \end{eqnarray} where $\psi(m)$ is the Euler's psi function. The expansion (\ref{kerco1}) is different from the behavior of the non-relativistic potentials. However, a procedure similar to that of the non-relativistic case can be applied to cancel the lowest order terms in the expansion (\ref{kerco1}) and generate a non-vanishing kernel in the limit $\mu \to 0$, since at large $R_E$ there is the exponential suppression similar to the one in the Yukawa potential: \begin{eqnarray} {\cal K}_{ladder} (R_E\to \infty,\mu) \sim g^2 \mu^{1/2} \frac{{\sf exp}[-\mu R_E]}{R_E^{3/2}} . \label{kerco2} \end{eqnarray} We can see that the direct use of operation (\ref{dyukp}) does not lead to the desired non-vanishing at $R_E\to\infty$ behavior of the kernel. We study here the kernel with the lowest power in $\mu^{-1}$, which as we expect controls the asymptotic behavior at large $R_E$. Analysis in the momentum space, similar to that of the non-relativistic case, gives the following conditions for the coefficients $C_j$ (see Appendix~B): \begin{eqnarray} C_j = \mu^{-2}\tilde C_j; \quad \sum_j \tilde C_j = 0; \quad \sum_j \alpha_j^2\tilde C_j = 0; \quad \sum_j \alpha_j^2{\sf ln}\alpha_j\tilde C_j \equiv A \ne 0, \label{prel} \end{eqnarray} which can be satisfied explicitly for $j_{max}=3$. Note, that the second condition cancels the most singular terms, $\sim R_E^{-2}$, in the coordinate space, (\ref{kerco1}), similar to the non-relativistic potentials. This choice of conditions leads to a kernel of the form: \begin{eqnarray} {\cal K}_{con}(k) = - {(2\pi)^4}U^4\delta^{(4)}(k_E), \label{krel} \end{eqnarray} where for simplicity we introduce new effective coupling constant $U$, which has dimension of mass, and sign is chosen to provide us with a confining-like kernel, similar to the positive constant potential in the non-relativistic case. Fourier transform of the kernel, (\ref{krel}), is a constant, $\sim U^4$, in four dimension. This means that it does not behave like a non-relativistic constant potential, for which we expect behavior like $\sim \delta(t_0) \cdot constant$. Therefore, the kernel (\ref{krel}) does not exactly correspond to the non-relativistic constant potential and the effective constant, $U^4$, is not related to constant in such a potential. Note, since the constant $U^4$ does not have the direct physical meaning, the choice of the factor ${(2\pi)^4}$ is arbitrary and it is made for further simplification of formulae. The form of the kernel, (\ref{krel}), in accordance with our main idea, is considered as a special limiting case of the sum of the ladder kernels (sum of one-boson exchanges), which provides explicit covariance of the kernel and connection with the usual field-theoretic constructions. (Note that this is a valid form in the Euclidian space, whereas the transition to the Minkowsky space is not defined.) By itself the $\delta$-form of the kernel is not something very unusual in studying of the bound states of the quarks. For instance, such a form is considered as "regularized" form of the highly singular, $\sim k^{-4}$, behavior of the gluon propagator~\cite{qpr} and is a basis of the models for studying of the SD equation for the quark propagator~\cite{munczek,williams}. In particular, this form of the gluon propagator leads to the quark propagator without singularities along physical momentum. In the lowest order, such a gluon propagator gives the kernel of the BS equation~\cite{munczek}. Let us study the effect of the kernel, (\ref{krel}), in the BS equation, (\ref{sbse}), under the Wick rotation. The form of ${\cal K}_{con}$ allows an integration in the equation explicitly: {\begin{eqnarray} \psi (p,P) = - \frac{{U^4}}{ D(p_1,p_2) } \psi(p,P). \label{1con} \end{eqnarray} } {}From eq.(\ref{1con}) we find that the kernel, ${\cal K}_{con}$, does not allow for the bound state solutions of the BS equation. In this sense we can expect that this kernel, in effect, is similar to the constant non-relativistic potential, which along does not allows for the bound states. If this is the case, we expect that adding this kernel to the regular attractive kernel, i.e. like the one presented in Section~2.3, Table~1, we will get a shift of the spectrum by a constant. The BS equation with the combined kernel, ${\cal K}_G+{\cal K}_{con}$, can be transformed to the usual form, but with a modified two-body propagator: {\begin{eqnarray} \psi (p,P) = \frac{1}{D(p_1,p_2) + U^4}\int \frac{d^4 p'}{(2\pi)^4}{\cal K}_G(p,p',P)\psi(p',P). \label{2con} \end{eqnarray}} New propagator in eq. (\ref{2con}) has different analytical properties, compared to the initial free propagator under the Wick rotation, (\ref{twopropr}), and main difference is that the new propagator does not have singularities at $M > 2m$, which were a signal of the limit of the spectrum at this point. This is an indication that the {\em physical spectrum} exists beyond this point. However, without numerical analysis we are not able to discuss the properties of the solutions at $M > 2m$. Before to go to the numerics, let us discuss the possible effects of the self energy corrections in the presence of the interaction of the form (\ref{krel}). Dealing with the kernel (\ref{krel}) as a sum of the lowest order (ladder) kernels, we calculate the one-loop self-energy corrections to the single quark propagator, $d(p)^{-1}$. Integration over the loop is performed at imaginary $p_0$ component of the four-momentum of the quark and result can be analytically continued to any values of $p_0$. For physical momentum we get: {\begin{eqnarray} d(p) = p^2 -m^2 +\frac{U^4}{p^2-m^2}. \label{singprop} \end{eqnarray}} The propagator (\ref{singprop}) does not have singularities for physical values of the momentum, $p$. This property of the interaction of the form (\ref{krel}) has already been established within model investigations of the behavior of the quark propagator~\cite{williams,qpr}. Now we see that this single quark effect is in a generic relation with the two-body confining interaction in the framework of the BS equation. For our calculations it is also important that the singularities of the modified two-body propagator in (\ref{2con}) with $D(p_1,p_2)=d(p_{1})\cdot d(p_{2})$ defined by eq. (\ref{singprop}) still allow to perform the Wick rotation. \subsection{Numerical investigation of the Bethe-Salpeter equation with the confining kernel} First we compare the meson spectra obtained with eq.~(\ref{2con}) with $U^4 = 0$ and $U^4 \ne 0$. For convenience we take the same kernel, ${\cal K}_G$, as that in the section~2.3. The main effect, we expect here, is the shift of the spectrum beyond the limiting point $M_{lim} = 2m$. In order to compare the spectra, we calculate $\lambda(M)$ for two lowest states in the channels with $L=0,1$. This is enough to make conclusions about (i) the position of the bound states, (ii) limiting point of the spectra and (iii) separation between ground and excited states. Results of these calculations are presented in the Fig.~3, groups of curves A and B. The constant $U$ for the case B is chosen to be of the typical energy scale of the equation, $U=m$. We find, indeed, that the masen spectrum of the equation with $U\ne 0$ is extended beyond the point $M_{lim}$. However, it displays unusual behavior beyond the point, $M_{lim}^B$ (this point is shown on Fig.~3 by the arrow) \cite{foot3}. Non-monotonical behavior of the curves $\lambda(M)$ indicates some difficulties in the interpretation of the corresponding solutions. The obvious difficulty is the existence of two solutions with the same coupling constant $\lambda$ and different masses, $M$, (see e.g. discussions in refs.~\cite{scale}). Therefore, we have to find a way to isolate the only one physical solutions. The solution of this problem is quite simple. Calculating the normalization of the solutions of the equation along $M$ we find that the ground state solutions beyond $M_{lim}^B$ have a negative norm, i.e. are the {\em abnormal} non-physical solutions. \cite{foot4}. This observation gives us the selection rule to eliminate the extra solutions. Another problem with the solutions corresponding $U^4 \ne 0$ is that spectrum is not only shifted to larger masses, but also the separation between levels is drastically increased (see Table~II). This effect make it a problem for a phenomenological use of the kernel of the form (\ref{krel}). Indeed, if we intend to consider states above $M_{lim}$ we have to take $U^4$ large enough to provide us with a new limit of the spectrum, however this can be in conflict with the desired separation between levels. One may try to adjust the parameters of the remaining part of the full kernel, ${\cal K}_G$, so as to have a reasonable density of levels. However our analysis showed that in the presence of the kernel, ${\cal K}_{con}$, separation between levels depends only weakly on the kernel, ${\cal K}_G$. \vskip 6mm \begin{center} \begin{tabular}{|c|c|c|c|c|} \hline $\Delta$ M & A & B & C & D\\ &$\lambda^A_{phys}=0.13$&$\lambda^B_{phys}=0.25$ &$\lambda^C_{phys}=0.44$&$\lambda^D_{phys}=1.44$\\ \hline \hline 1P - 1S & 0.49~GeV & 0.77 ~GeV & 0.48~GeV & 0.48~GeV \\ \hline 2S - 1P & 0.19~GeV & 0.47 ~GeV & 0.16~GeV & 0.17~GeV \\ \hline 2P - 2S & 0.21~GeV & 0.42 ~GeV & 0.22~GeV & 0.26~GeV \\ \hline \end{tabular} \vspace*{5mm} Table. II. The levels splitting in the "scalar charmonium" for different kernels . \end{center} It should be remembered that in the BS equation the addition of interaction of any kind to the full kernel cannot be related linearly to the shift in the mass of the system. The same interaction also "shifts" the masses of the constituents through the self-energy corrections to the single-particle propagators. A self-consitent approach has to be adopted to include both the self-energy corrections and changes in the two-body interactions. We take into account the self-energy correction, (\ref{singprop}), to the quark propagator. This leads to a corresponding modification of the two-body propagator $D(p,P)^{-1}$, (\ref{twoprop}). The resulting spectrum is presented in Fig.~3, group of curves C. It is clear that for the interval of masses $\sim 2 \div 3.5$~GeV the density of levels "returns to normal", of the same value as in the case A (see Table~II). The value of constant $\lambda_{phys}^C = 0.44$ is chosen as an example giving the spectrum density close to the the one of the case A. At smaller $\lambda_{phys}^C $ the separation of levels is even smaller. The examples of calculations, B and C, show that the kernel containing the part ${\cal K}_{con}$ is indeed similar in its effect to the non-relativistic potential in coordinate space, $V(r)$, with $V(r\to \infty) \to V_{\infty}$. where the positive real constant $V_{\infty}$ defines the shift of the bound state spectrum compared to the case with no ${\cal K}_{con}$, the case A. From Fig.~3 another similarity with the non-relativistic constant potential is obvious, this kernel gives only a limited number of states in the spectrum, which can be adjusted by varying the $\lambda_{phys}$. For instance, for the case C with $\lambda^C_{phys}=0.2 $ there is only one bound state, $1S$, whereas with $\lambda^C_{phys}=0.45$ there are four states in the S and P channels (there can be other undetected states in channels with higher $L=2,\ldots$). On the other hand, taking account of the self-energy corrections, corresponding to this type of interaction, leads to the disappearance of the poles in the single quark propagator. Therefore quarks cannot propagate other than being bound in a bound state. Apparently, this fact is not related to the infinitely rising interaction between them, but rather to the modification of the single quark properties. To be sure that the picture is valid in a wide interval of the constant $U^4$, also the spectrum is calculated for the case $U^4 = 4m^4$ and self-energy corrections taken into account. The result is shown in the Fig.~3, group of curves D. That the picture is found to be similar to the one of case C, but the spectrum is shifted up to even larger $M_{lim}^D\sim 4$~GeV. The density of levels in the interval of masses $\sim 2.5 \div 4$~GeV is the same as those of cases A and C (see Table~II). The value of constant $\lambda_{phys}^C = 0.44$ is chosen again to give the spectrum density close to the the one of the case A. Note that in our model the constant $U$ can not be taken arbitrary large, since all the calculations are performed in the lowest order. The natural criteria on the maximum value of $U$ is that the corrections of the lowest order, $\sim U^4$, must not be too large. For instance, a shift of the mass spectrum with $U\ne 0$ should not be too big compare to the typical masses for the case with $U=0$. \section {Conclusions} We have presented a method to find the ground state and excited states of the two-body Bethe-Salpeter equation for a channel with any quantum numbers. This method allows us to solve the bound state problem without reduction of equation to quasipotential form or any other approximations. Based on a qualitative analogy with the construction of a non-relativistic potential with non-vanishing asymptote at large distances, $r\to\infty$, we have proposed a recipe to obtain the confining kernel for the Bethe-Salpeter equation, parametrized in the form of a special limiting case of a superposition of ladder kernels. We find that in the simplest case such a kernel is proportional to $\delta(k_E)$ in the Euclidean momentum space, which corresponds to a constant kernel in the coordinate space. We have studied the effect of this kernel on the spectrum of the Bethe-Salpeter equation, when the usual attractive interaction is added. It is found that this kernel is similar in its effect to the non-relativistic potential in coordinate space, $V(r)$, with $V(r\to \infty) \to V_{\infty}$. The positive real constant $V_{\infty}$ gives the scale that defines the limit of the bound state spectrum compared to the sum of the constituent masses, $M < 2m + V_{\infty}$. At the same time the self-energy corrections remove the singularities from the propagators of the constituents, i.e. constituents do not propagate. Combination of these features of the solutions allows an interpretation of this type of interaction as a confining interaction. The illustrative analytical and numerical calculation are presented for a model of massive scalar particles with scalar interaction and do not pretend to a phenomenological application. However, the developed formalism can be straightforwarly adopted for the Bethe-Salpeter equation for the bound state of two spinor fields and, therefore, can be used for the realistic studies of the properties of the quark-antiquark systems, mesons. \section{Acknowledgments} The authors thank A. Maximov and Yousuf Musakhanov for useful discussions. One of us (A.U.) is grateful to F. Gross, J. Milana and A. Stadler for the interesting discussions and comments. The research is supported in part by the Natural Sciences and Engineering Research Council of Canada. \vspace*{.5cm} { \Large \bf Appendix A. The non-relativistic confining potentials in the momentum space.} \setcounter{equation}{0} \defB.\arabic{equation}{A.\arabic{equation}} \noindent In order to establish the form of the potentials defining by the eq. (\ref{syukr}), (\ref{p0}) and (\ref{p1}), let us consider auxiliary integral, $I_a$: \begin{eqnarray} I_a &=& \int \frac{d^3 {\bf k}}{(2\pi)^3} V_a (k) f({\bf k}) \label{b1}\\ &=&\int \frac{d\Omega}{(2\pi)^3} \int\limits_0^{\infty} dk k^2 V_a (k) f(k, \Omega), \label{b11} \end{eqnarray} where $f({\bf k})$ is any function for which we assume a "good" behavior ($f \to 0$ when $k\to \infty$, no singularities, existence of derivatives, etc.), $a = c, l$ depending on which of conditions the (\ref{p0}) or (\ref{p1}) is imposed on the potential and $V_a(k)$ defined as a fourier transform of eq. (\ref{syukr}): \begin{eqnarray} V_a (k) &=& 4\pi\sum_j \frac{4\pi C_j}{k^2 + \alpha_j^2 \mu^2}. \label{b2} \end{eqnarray} The limit $\mu \to 0$ is assumed and we take it later. Using the common condition $\sum C_j =0$, we rewrite eq. (\ref{b11}) as \begin{eqnarray} I_a &=& -4\pi\mu^2\int \frac{d\Omega}{(2\pi)^3} \sum_j C_j\alpha_j^2 \int\limits_0^{\infty}\frac{ dk}{k^2 + \alpha_j^2 \mu^2} f(k, \Omega). \label{b3} \end{eqnarray} Integrating by parts the last integral in (\ref{b3}), we get: \begin{eqnarray} I_a &=& 4\pi\mu \int \frac{d\Omega}{(2\pi)^3}\sum_j C_j\alpha_j \int\limits_0^{\infty} dk \;{\sf arctg}\left [ \frac{k}{\alpha_j \mu} \right ] f'(k, \Omega), \label{b4} \end{eqnarray} where $f'= \partial f /\partial k$. Let us now consider integration over $k$ only. These integrals on the r.h.s. of eq. (\ref{b4}) can be split in two parts: \begin{eqnarray} I_a \propto \mu \sum_j C_j\alpha_j \left \{ \int\limits_0^{\alpha_j\mu} + \int\limits_{\alpha_j\mu}^{\infty} \right \} dk \;{\sf arctg}\left [ \frac{k}{\alpha_j \mu} \right ] f'(k, \Omega). \label{b5} \end{eqnarray} We estimate the first integral by the mean value theorem: \begin{eqnarray} \mu \sum_j C_j\alpha_j \int\limits_0^{\alpha_j\mu} dk \ldots = \mu^2 \left( \frac{\pi}{4}-\frac{{\sf ln}2}{2} \right )\sum_j C_j\alpha_j^2 f'(\xi\alpha_j\mu, \Omega), \label{b6} \end{eqnarray} where $0\le \xi \le 1$. The second integral is estimated using the expansion ${\sf arctg}x = \pi/2 -1/x+1/(3x^3)-\ldots$, which is valid at $x\ge 1$. It can be shown that the first term in this expansion gives the leading contribution to the full integral in the limit $\mu \to 0$: \begin{eqnarray} \mu \sum_j C_j\alpha_j \int\limits_{\alpha_j\mu}^{\infty} dk \ldots = -\frac{\pi}{2}\mu \sum_j C_j\alpha_j f(\alpha_j\mu, \Omega). \label{b7} \end{eqnarray} Finally, taking the limit $\mu \to 0$ and accounting for the conditions on the coefficients $C_j$, we find \begin{eqnarray} I_c &=& A_c f({\bf 0}), \label{b8}\\ I_l &=& \left( \frac{1}{2}-\frac{{\sf ln}2}{\pi} \right ) A_l f'({\bf 0}). \label{b9} \end{eqnarray} These equations give us the potentials in the form (\ref{p0p}) and (\ref{p1p}). { \Large \bf Appendix B. The relativistic confining kernel in the momentum space.} \setcounter{equation}{0} \defB.\arabic{equation}{B.\arabic{equation}} \noindent In order to establish the form of the kernel defined by eq. (\ref{conk}) at the limit $\mu \to 0$ and lowest (but non-zero) degree of $\mu^{-1}$ in the $C_j$, let us consider the auxiliary integral, $I$: \begin{eqnarray} I &=& \int \frac{d^4 {\underline k}}{(2\pi)^4} {\cal K}_{con}(k) f({\underline k}) \label{c1}\\ &=&\int \frac{d\Omega^{(4)}}{(2\pi)^4} \int\limits_0^{\infty} dk k^3 {\cal K}_{con} (k) f(k, \Omega^{(4)}), \label{c11} \end{eqnarray} where $\underline k$ is four-momentum in Euclidian space, $k = ({\bf k}^2 + k_0^2)^{1/2}$, $\Omega^{(4)} $ is the hyperangle defining the orientation of the vector $\underline k$ in the four dimensional space; $f({k})$ is an arbitrary function for which we assume "good" behavior ($f \to 0$ when $k\to \infty$, no singularities, existence of derivatives, etc.). The limit $\mu \to 0$ is assumed and we take it later. We are omitting all nonessential factors, such as $2\pi$, etc., in the following calculations. Adding and subtracting expression $C_j/k^2$ to each term in (\ref{conk}), we rewrite eq. (\ref{c11}) as \begin{eqnarray} I \propto && -\mu^2\int {d\Omega^{(4)}}\sum_j C_j\alpha_j^2 \int\limits_0^{\infty}\frac{ dk \, k}{k^2 + \alpha_j^2 \mu^2} f(k, \Omega^{(4)}) \label{c3}\\ &&+ \int {d\Omega^{(4)}}\sum_j C_j \int\limits_0^{\infty}\frac{ dk}{k^2} f(k, \Omega^{(4)}). \nonumber \end{eqnarray} Th last term is cancelled by the condition \begin{eqnarray} \sum_j C_j = 0. \label{cc1} \end{eqnarray} Let us now consider integration over $k$ only. This integral in r.h.s. of eq. (\ref{c3}) can be evaluated as: \begin{eqnarray} I \propto -\mu^2 \sum_j C_j\alpha_j^2 f(\alpha_j\mu, \Omega^{(4)}) \int\limits_0^{\Lambda} \frac{dk \; k}{k^2+\alpha_j^2 \mu^2}, \label{c5} \end{eqnarray} since the function remaining under integration has sharp maximum at $k=\alpha_j\mu$. We also introduce the cut-off parameter $\Lambda$ to regularize formally the logarithmically divergent integrals. At a later stage of calculation we take the limit $\Lambda \to \infty$, but for the moment it is enough to assume $\Lambda \gg \mu$. Performing the integration we get \begin{eqnarray} I \propto -\mu^2 \sum_j C_j\alpha_j^2 f(\alpha_j\mu, \Omega^{(4)}) \left [ {\sf ln}\left ( \frac{\Lambda}{\mu} \right ) -{\sf ln}\alpha_j \right ]. \label{c6} \end{eqnarray} Applying the condition \begin{eqnarray} \sum_j C_j\alpha_j^2 = 0,\quad \sum_j C_j\alpha_j^2{\sf ln}\alpha_j \ne 0, \label{cc2} \end{eqnarray} and taking the limit $\mu \to 0$, we get \begin{eqnarray} I \propto \mu^2 f(0)\sum_j C_j\alpha_j^2 {\sf ln}\alpha_j, \label{c7} \end{eqnarray} which proves the form of the kernel as eq. (\ref{krel}).
\section{Introduction} In spite of the large amount of research done on C$_{60}$ and its compounds since its discovery \cite{kroto}, the electronic structure and the importance of solid state band structure effects remain controversial. Solid C$_{60}$ seems to exhibit a dualistic behaviour. On one hand it behaves like a molecular crystal in which the molecular properties (like the vibrational states and electronic excitations) are only weakly perturbed by the crystal symmetry, but on the other hand it behaves like a semiconductor, with a moderate (2.3 - 2.6 eV \cite{lof,weaver}) band gap, which can be electron doped resulting in low energy impurity states and band widths of about 0.6 eV \cite{saito,shirley}. Also quite different from typical molecular crystals is that C$_{60}$ forms ionic compounds, which in some cases exhibit metallic and even superconducting \cite{hebert} behaviour, clearly demonstrating the importance of one-electron band formation. With regard to the properties of these compounds they are reminiscent of the charge transfer type of molecular solids like the much studied TCNQ salts except that the C$_{60}$ compounds usually show 3-dimensional behaviour rather than the 1- or 2-dimensional behaviour exhibited by the charge transfer molecular solids. This, obviously, is due to the spherical rather than linear or planar structure of the molecule. In this paper we present evidence that in pure C$_{60}$ also the excitons exhibit this dualistic behaviour. The energies of the excitonic states are close to those of the gas phase molecule emphasizing the molecular characteristics, but the propagation of the exciton results in abnormally large excitonic band widths and mixings of different multiplets for a molecular solid. This can be explained within a one- and two-particle band structure theory. It is well established that the strong delocalization of the p$_{\pi}$-electron network (as is also the case on a C$_{60}$ molecule) can result in strong nonlinear optical effects \cite{poly-nlo}. Koopmans {\em et al.} \cite{koopmans-b,koopmans-res} have shown that the SH signal is very strong due to a double resonance if the primary energy is tuned to the $^1$T$_{1g}$ excitonic state at 1.81 eV. This provides a possibility to study the excitons inside the electronic band gap, in particular the exciton band width, the band splitting due to crystallographic phase transitions, and the mixing of multiplets due to the crystal symmetry. To facilitate this study we have developed a theory for the exciton splittings and dispersions based on the molecular multiplet splittings and solid state effects arising from a charge transfer mechanism for the exciton propagation \cite{lof,robert}. The one-electron (-hole) hopping integrals required for this are obtained from a tight-binding fit to the LDA band structure of C$_{60}$ as given by Satpathy {\em et al.} \cite{satpathy}. We show that reasonable agreement can be obtained with the experimental SH line shape with only one adjustable parameter, namely the $^1$T$_{1g}$ - $^1$G$_g$ molecular multiplet splitting. \section{Experimental} C$_{60}$ with a purity better than 99.99\% was evaporated from a Knudsen cell onto a substrate at UHV pressures below 4$\cdot$10$^{-9}$ mbar. As substrates we used fused quartz or at low temperature MgO, a good thermal conductor. For the SH experiments a Nd:YAG laser was used to pump a dye laser, producing 7ns pulses with an energy of approximately 6 mJ/pulse and a repetition rate of 10 Hz. The fundamental frequency was scanned in the range $\hbar \omega$= 1.7...2.0 eV. The SH intensities were calibrated by using a reference quartz crystal in a transmission geometry, carefully tuned to a Maker fringe optimum by fine-tuning the frequency, and corrected for changes of the coherence length in the quartz crystal as function of the photon energy. All SH experiments are performed at a fixed angle of incidence (45$^{\circ}$ to the surface normal) and the specular reflected SH signal was detected. The SHG of thin C$_{60}$ films exhibit complicated thickness- (and through the dispersion also frequency-) dependent interference phenomena. We showed before \cite{koopmans-b} that for a mixed$_{\mbox{in}}-$p$_{\mbox{out}}$ polarization (m-p) combination (mixed means 50\% p and 50\% s polarized light) the SH interference pattern exhibits a broad minimum for C$_{60}$ film thicknesses of around 250 nm. Therefore we chose a m-p polarization combination and a thickness of 250 nm for the measurements presented in this paper, so that dispersive interference effects can be neglected. The temperature dependent SHG experiments were performed using a He-flow cryostat (4 - 500 K). The temperature was measured with a thermocouple glued to the substrate. Possible effects of heating during the laser pulse were examined by varying the laser power. We found that below 100 K the temperature during the laser pulse was about 20-30 K higher than the one measured with the thermocouple. At higher temperatures, in particular around the rotational ordering phase transition temperature (260 K), no heating by the laser pulse was detected. \section{Experimental results} In Fig.1 the SH intensity measured at various temperatures is shown as a function of the fundamental photon frequency ($\hbar \omega$). Around room temperature we observe the resonance at about 1.8 eV already previously reported \cite{koopmans-b,koopmans-res}. Notice for decreasing temperature the strong enhancement of the SH intensity, the overall blue shift and the temperature dependence in the line shape of the resonance. Fig.2a and 2b show the temperature dependence of the zeroth and first moment of the spectrum corresponding to the integrated intensity and to the mean frequency, respectively. In both cases we see a strong temperature dependence at the phase transition temperature of 260 K. It is also just below this temperature that we observe a splitting of the resonance into two peaks with an intensity ratio of about 3:1. The splitting is about 40 meV. The total width of the signal at the base of the line is approximately 100 meV, which is very large for an exciton band width of a molecular crystal, as discussed below. \section{Discussion} Looking at the data in Fig.1 and 2, there are three main features to be explained: $(i)$ the splitting of the signal below the phase transition; $(ii)$ the line width larger than expected for an electric dipole forbidden transition in a molecular crystal, and $(iii)$ the overall strong temperature dependence of the SH intensity and line shape. We propose three possible mechanism for the splitting, namely a Jahn-Teller effect, a Davydov splitting and a mixing of the electronic molecular states. A {\em Jahn-Teller splitting} can probably be discarded since it is expected to be at most 12 meV for the singlet states as determined by Wang {\em et al.} \cite{yu}. Before discussing the other two possibilities we briefly review the basic ideas involving the propagation of excitations and their observation by SHG. For a schematic picture of the double resonant SHG process observed in this energy range we refer to Fig.1 of ref \cite{koopmans-b}. The three level diagram consists of a magnetic dipole transition from the molecular ground state to the $^1$T$_{1g}$ excited state (involving a h$_{u} \rightarrow$ t$_{1u}$ single-electron transition), followed by an electric dipole transition to a $^1$T$_{1u}$ state at about 3.6 eV (h$_g \rightarrow$ h$_u$), and finally an electric dipole transition back to the ground state (t$_{1u} \rightarrow$ h$_g$). Linear optical experiments exhibit a strong electric dipole allowed transition at 3.56 eV with a half width at half maximum of 0.23 eV. Since this width is much larger than the one observed in our experiment (0.06 eV) we concluded that the sharp features in the SHG spectrum must be related to the intermediate $^1$T$_{1g}$ exciton state \cite{k-thesis,k-saltlake}. This difference in width can easily be understood by comparing the intramolecular excitations, where the electron and the hole are bound, with the intermolecular electron-hole excitations. The latter determine the conductivity gap involving dissociated electron-hole states. As measured by photoconductivity \cite{photocond}, or by combined photoelectron and inverse-photoelectron spectroscopy \cite{lof} this gap is 2.3 eV. In Fig.3 we show the energy level scheme of the intramolecular excitonic excitations (on the left hand side) and the solid state intermolecular band gap excitations (on the right hand side). The molecular $^1$T$_{1u}$ state at 3.6 eV is well inside the intermolecular electron-hole continuum and will decay into this with a hopping integral comparable to the one-electron (-hole) band width. The $^1$T$_{1g}$, however, is an electron-hole bound state, inside the band gap, and will therefore have a long lifetime. The extra energy required to dissociate the electron-hole pair of the exciton (i.e. the exciton binding energy) is directly related to the onsite Coulomb interaction measured to be about 1.3 - 1.6 eV by Lof {\em et al.} \cite{lof}. Since the $^1$T$_{1g}$ exciton is bound and the transition to the ground state is electric dipole forbidden, we expect it to be very long lived and we would expect a very small exciton dispersion by conventional optical dipole - optical dipole intermolecular propagation. The still quite large total width of more than 100 meV is therefore difficult to understand in the limit of a molecular solid. It is, however, well known that band structure effects in C$_{60}$ are not neglegible. Lof {\em et al.} \cite{lof} already suggested a propagation mechanism which could lead to a substantial exciton dispersion. A similar mechanism involving virtual charge transfer states was previously suggested by Choi {\em et al.} \cite{silbey} to explain the dispersional width of optically forbidden excitons in molecular crystals. This propagation mechanism is shown pictorially in Fig.4 which demonstrates how an electron-hole pair on site $i$ can propagate to a neighbouring site $j$ via a virtual excited intermediate state. This intermediate nearest-neighbour (charge transfer) state, in which the electron is on site $i$ and the hole on a nearest-neighbour site $j$ (or vice versa), is at an energy $U - V$ (the difference between the onsite Coulomb interaction and the nearest-neighbour Coulomb repulsion \cite{bluhwiler,comment-m}) higher than the exciton ground state energy. The net effective exciton hopping integral is given from perturbation theory by: \begin{equation} T^{\mbox{exciton}} = \frac{2 t_e t_h}{U-V} \end{equation} where $t_{e(h)}$ are the average single electron (hole) nearest-neighbour hopping integrals. It should be noticed that the same electron and hole hopping integrals are involved for the red shift of an exciton energy in the solid relative to the corresponding one in the gas phase. The red shift is then given in first perturbation theory by: \begin{equation} \Delta E = K (\frac{t^2_e + t^2_h}{U-V}) \end{equation} where $K$ is a geometrical factor related to the symmetry of orbitals and the nearest-neighbour coordination number. As an example, one has $K$ = 12 \cite{auger} for a totally symmetrical one-electron and one-hole orbital in a FCC lattice. To estimate the red shift of the $^1$T$_{1g}$ state we look at the isolated C$_{60}$ molecule where Gasyna {\em et al.} \cite{gasyna} found the $^1$T$_{1g}$ at about 1.92 eV. The same value has been assigned to the $^1$T$_{1g}$ state by Negri {\em et al.} \cite{negri} using the absorption spectra of C$_{60}$ in n-hexane solution of Leach {\em et al.} \cite{leach}. Koopmans {\em et al.} found the $^1$T$_{1g}$ in solid C$_{60}$ resonant at 1.81 eV. This means that in the solid state this exciton is red shifted by 110 meV, which is in the same order as the exciton band width. This supports the above mentioned exciton hopping mechanism. Detailed calculations of the one-electron (-hole) hopping integrals (discussed further on) show that only nearest-neighbour hopping have to be considered \cite{robert}. Therefore, this exciton hopping mechanism does not destroy the molecular Frenkel character of the exciton. The above described mechanism for the dispersion of a Frenkel exciton is analogous to the charge transfer mechanism proposed by Lof {\em et al.} \cite{lof} and to the mechanism originally used to describe excitons in molecular charge transfer salts \cite{silbey}. It is also very similar to the so called superexchange mechanism used to describe Frenkel d-d excitons in 3d transition metal compounds \cite{allen,sugano}. Assuming that the exciton band width is primarily due to such a dispersional width we look again at the temperature dependence of the line shape. First of all we might have expected to see only the zero quasi momentum ({\bf k} = 0) exciton because of the long optical wavelength. At high temperatures, however, the molecules are rapidly rotating resulting in dynamic orientational disorder which will cause a break down of the translational lattice symmetry and of the $\Delta${\bf k} = 0 selection rule. In the extreme case we would expect to see just the total exciton density of states as we believe is indeed the case at high temperatures. Upon lowering the temperature below the phase transition at 260 K the rotations are strongly reduced, leading to a decrease of dynamic disorder and therefore to approach the $\Delta${\bf k} = 0 selection rule. Since {\bf k} then becomes a good quantum number we will see only the exciton states with {\bf k} vectors close to $\Gamma$, the center of the Brillouin Zone. In the low temperature phase there are four molecules per unit cell so that the exciton band at the $\Gamma$ point can split up into two or more bands. This splitting, called the {\em Davydov splitting}, which represents a first possible explanation, will be of the order of the exciton band width and is prominently due to the dependence of the exciton transfer integral on the relative orientation of neighbouring molecules. Recent two-photon excitation of C$_{60}$ single crystal at 4 K by Muccini {\em et al.} \cite{muccini} shows a band at 1.846 eV which is assigned to the same lowest forbidden Frenkel exciton of $^1$T$_{1g}$ symmetry as discussed in this paper. They also find a second band at higher energy (1.873 eV). They discuss this second band in terms of a crystal field effect and as a possible Davydov splitting. They give an alternative assignment of the second band as being due to a second forbidden electronic state. Indeed semi-empirical quantum-chemical calculations \cite{negri} have shown that there are several closely spaced forbidden states which lie in a narrow energy range \cite{muccini}. The two-photon spectrum of Muccini {\em et al.} strongly resembles the low temperature SH resonance in Fig.1. However, their two-photon absorption, being a third-order nonlinear optical experiment, involves other selection rules than our SHG experiment. In order to get a more detailed understanding of the results, we carried out the full exciton calculation starting from the basic ideas described by equation (1) and (2). The details will be published elsewhere \cite{robert}. Here we restrict ourselve to briefly describe the ingredients of the calculations and the results. In the full calculation the orbital degeneracy of the t$_{1u}$ (3 fold) and h$_u$ (5 fold) must be taken into account so that there are several electron and hole hopping integrals depending on the orbital quantum numbers. Satpathy {\em et al.} \cite{satpathy} have described how those can be obtained from one-particle band structure calculation using a tight-binding fit. The electron and hole hopping integrals are a function of the relative orientation of the buckyballs. These integrals are completely determined from a single fit to the band structure for a particular given structure. Also we must take into account the multiplet structure of the molecular excitations due to the intramolecular Coulomb interaction as described by Negri {\em et al.}. These multiplet splittings are not very well known but can be obtained from optical or electron energy loss data of the gas phase or in solution. The effective exciton transfer integrals are then a sum of products of electron and hole transfer integrals divided by $U - V$. The degree to which each of the electron and hole hopping integrals contribute to the dispersion of a particular exciton is determined by the weight of the electron-hole product function in the particular excitonic state under consideration. In addition to the broadening of the molecular multiplets into bands, there is also a mixing of the various molecular exciton states because of the lowering of symmetry in the crystal. The only remaining parameters are $U - V$ and the molecular multiplet splitting. Conserning $U - V$ a rough estimate can be taken from the Auger data of Br\"{u}hwiler {\em et al.} \cite{bluhwiler}: U $\simeq$ 1.1 eV (0.2 eV is substracted because of the higher exciton energy of a singlet), and V $\simeq$ 0.7 eV. This leaves us with $U - V \simeq 0.35$ $\pm$ 0.2 eV. An independent estimate for $U - V$ can also be obtained from the experimental red shift as given in eq.(2). Taking $U - V$ = 0.35 eV we get a calculated red shift comparable to the experimental observed one. Concerning the multiplet splitting we will see below that all we need for the present propose is a small $^1$T$_{1g}$ - $^1$G$_g$ splitting. These exciton dispersion calculations, show that the $^1$T$_{1g}$ band at the $\Gamma$ point splits up in three $^1$T$_g$ bands, one $^1$A$_g$ and one $^1$E$_g$ band (Fig.5a) \cite{robert}. This can be expected from group theoretical arguments because of the transition from the space group Fm$\bar{3}$m of the high temperature phase to the space group Pa$\bar{3}$ of the low temperature phase \cite{harris,dresselh}. The Davydov splitting is found to be about 30 meV, which is close to the experimental splitting (40 meV) of the two peaks. The calculation, however, predicts that more than 90\% of the weight would be in the lowest $^1$T$_{g}$ band. This is inconsistent with our data! Another possible explanation appears when all molecular multiplet states and their mixing is included. As already mentioned, the quantum-chemical calculations of Negri {\em et al.} \cite{negri} show that the $^1$T$_{2g}$ and $^1$G$_g$ states (in terms of states of isolated C$_{60}$ molecules with icosahedral symmetry) are nearly degenerated with the $^1$T$_{1g}$. In the crystal, however, the point group symmetry is lower. This gives rise to a mixing of the icosahedral electronic eigenstates (compare with Table VIII of \cite{dresselh}). When this {\em mixing} of the $^1$T$_{1g}$ and $^1$G$_g$ Bloch states is taken into account in the exciton dispersion calculations \cite{robert}, a second somewhat smaller peak arises at higher energy. This is another possible explanation for the splitting. Accordingly, the main peak at 1.826 eV is (in terms of molecular states) a mixed state of $^1$T$_{1g}$ with some $^1$G$_g$ character, and the second peak at 1.866 eV is a $^1$G$_g$ state with some $^1$T$_{1g}$ character. Since in our SHG experiment we probe the magnetic dipole allowed transitions \cite{note}, only the $^1$T$_{1g}$ component is visible. This would explain the difference in intensity. Fig.5 shows the calculated spectrum with and without mixing in of the $^1$G$_g$ state. The "mixed" curve agrees well with the experimental data (at the lowest temperature). Although the $^1$T$_{2g}$ state is also very close to the $^1$T$_{1g}$, the calculations show that these do not mix, because the neighbouring molecules do not have the required orientation for allowing a mixing of the corresponding electronic orbitals. Notice that for the $^1$T$_{1g}$ and $^1$G$_g$ mixing the same exciton transfer integrals are involved as in the case of a Davydov splitting. Because of these exciton hopping integrals, described by eq.(1), such a large Davydov splitting (compared to common molecular crystals where an electric dipole forbidden transition is considered) and a $^1$T$_{1g}$ and $^1$G$_g$ mixing are possible. The blue shift of the first moment of the SHG spectrum (Fig.2b) for decreasing temperature most probably has its origin in the orientational ordering, which takes place at the phase transition temperature (260 K). In the low temperature phase (T $<$ 260 K) the C$_{60}$ molecules can only jump between two equilibrium positions and at T $<$ 100 K they are practically frozen in, whereas in the high temperature phase (T $>$ 260 K) the buckyballs rotate freely in all directions \cite{heiney,david}. Calculations of the electron (hole) transfer integrals for both phases show that the hopping integrals for the high temperature phase are larger than those for the low temperature phase. This means that in the low temperature phase, where a double carbon-carbon bond faces a pentagon or hexagon, the exciton propagation is less favourable, resulting in a narrowing of the band. Since we are probing the $^1$T$_{g}$ state, which forms the bottom of the band, a narrowing of the band gives rise to a blue shift of the $^1$T$_{g}$ state. The difference in magnitude of the low and high temperature hopping integrals has its impact on still another process. Eq.(2) gives the relation between the electron (hole) hopping integrals and the red shift of a state in the solid compared to the gas phase. Thus, we expect that at low temperature (where the hopping integrals are smaller than those for the high temperature phase), the $^1$T$_{1g}$ state is less red shifted than at high temperature. This also resultes in a blue shift. Calculations, however, show that the first process will be dominant in our observed blue shift. What about the strong temperature dependence of the SH intensity? Also this can be explained in terms of the dynamic rotational disorder. SHG depends strongly on the retention of coherence in the $^1$T$_{1g}$ intermediate exciton state in a time scale determined by the excitation transition matrix elements. The intensity of the SHG goes like (N$_{\mbox{coh}})^2$ where N$_{\mbox{coh}}$ is the number of molecules coherently (in phase) contributing to the signal. Rotational motion during excitation results in dephasing of true oscillators so that N$_{\mbox{coh}}$ is expected to decrease strongly with temperature. This is a so called T$_2$ (i.e. dephasing time) like relaxation process. One expects, therefore, that the rotation time of a buckyball is of the same order as the time between the first and second transition. We estimate the time of revolution for a freely rotating buckyball at temperature T as: $\tau_{rot} = 2 \pi \sqrt{I / 2k_BT}$, where $I$=1$\cdot$10$^{-43}$ kgm$^2$ is the moment of inertia of the buckyball. For the time between the magnetic dipole transition and the first electric dipole transition by Fermi's Golden Rule gives $\tau^{-1} \simeq r^2_{ball} \cdot G \cdot \rho_{final}$, where $r^2_{ball}$ is the radius of the buckyball, $G$ is the energy current of the laser pulse and $\rho_{final}$ is the DOS of the final states (with one electron in the t$_{1u}$ (LUMO) and one hole in the h$_g$ (HOMO $-$ 1), the second highest occupied molecular orbital). We choose $\rho_{final}$ = W$^{-1}$ where W$\sim$0.5 eV is the band width of LUMO or HOMO $-$ 1. At T = 300 K we find $\tau_{rot}\simeq \tau\simeq10^{-11}$ s. This means that, in the time spent between the first and second transition a given buckyball can well performe a full rotation. Since the rotation of the different molecules is uncorrelated, this leads to strong (T$_2$) dephasing and decreasing of the SH intensity. Also a strong temperature dependence of the intensity can be found in photoluminescence experiments. There, the increase in intensity occurs at a somewhat lower temperature compared with the phase transition and the interpretation is still controversial \cite{pichler,graham,sauvajol,matus,schlaich}. Liu {\em et al.} have also done temperature dependent SHG of C$_{60}$ films, and they also observe a jump in the SH intensity around the orientational phase transition \cite{temp-china}. Their increase of the SH intensity for decreasing temperature is much less than what we observe. That is probably because they have done the SHG experiment using a fixed frequency (1064 nm) which involves other transitions well off the double resonance as found by Wilk {\em et al.} \cite{dieter}. Their SH intensity might also contain a change due to a shift of the exciton states on going through the phase transition temperature of 260 K as we have described above. \section{Conclusions} We have studied the dynamics of the $^1$T$_{1g}$ Frenkel exciton at $\hbar\omega$ = 1.81 eV with temperature dependent SHG. We find a very strong temperature dependence of this double resonance. Its SH intensity increases strongly close to the phase transition down to about 200 K. We explain this by correlating the rotational disorder of the C$_{60}$ molecules to a T$_2$ dephasing mechanism. Below the rotational ordering phase transition the SH resonance splits in two bands. Several ideas about what could be the cause of this splitting are discussed. Detailed exciton dispersion calculations taking into account the full symmetry, multiplet structure and crystal structure, yield large exciton dispersions with total band widths of about 100 meV and a Davydov splitting of the $^1$T$_{1g}$ state of 30 meV. The corresponding SH intensity is calculated to be concentrated to more than 90\% in the lowest Davydov component and this is not the observed behaviour. The experimental data including the two component structure in the low temperature phase is, however, very well described by the theory if the full multiplet structure and mixing of the $^1$T$_{1g}$ and $^1$G$_g$ states due to the lowering of space group symmetry is included. The experimental data together with the theory support strongly the ideas that the excitons in solid C$_{60}$ are Frenkel like but propagate via virtually excited charge transfer states described by Lof {\em et al.} \cite{lof}. This model is consistent with the total width, the splitting below the phase transition, the red shift relative to the gas phase and the blue shift with lowering temperature of the $^1$T$_{1g}$ exciton state component measured in our SHG experiment. \section {Acknowledgements} This investigation was supported by the Netherlands Foundation for Fundamental Research on Matter (FOM) with financial support from the Netherlands Organization for the Advancement of Pure Research (NWO).
\section{\@startsection {section}{1}{\z@}{-4.2ex plus -1ex minus -.2ex}{2.2ex plus .2ex}{\normalsize\bf}} \def\subsection{\@startsection{subsection}{2}{\z@}{-2.2ex plus -1ex minus -.2ex}{1.1ex plus .2ex}{\normalsize\bf}} \def\subsubsection{\@startsection{subsubsection}{3}{\z@}{-2.2ex plus -1ex minus -.2ex}{-1.2em}{\normalsize\it}} \def\@arabic\c@section.{\@arabic\c@section.} \def\thesection\@arabic\c@subsection.{\@arabic\c@section.\@arabic\c@subsection.} \def\thesubsection\@arabic\c@subsubsection.{\thesection\@arabic\c@subsection.\@arabic\c@subsubsection.} \def\@sect#1#2#3#4#5#6[#7]#8{\ifnum #2>\c@secnumdepth \def\@svsec{}\else \refstepcounter{#1}\edef\@svsec{\csname the#1\endcsname\hskip 1em }\fi \@tempskipa #5\relax \ifdim \@tempskipa>\z@ \begingroup #6\relax \@hangfrom{\hskip #3\relax\@svsec}{\interlinepenalty \@M \sec@upcase{#8}\par}% \endgroup \csname #1mark\endcsname{#7}\addcontentsline {toc}{#1}{\ifnum #2>\c@secnumdepth \else \protect\numberline{\csname the#1\endcsname}\fi #7}\else \def\@svsechd{#6\hskip #3\@svsec #8\csname #1mark\endcsname {#7}\addcontentsline {toc}{#1}{\ifnum #2>\c@secnumdepth \else \protect\numberline{\csname the#1\endcsname}\fi #7}}\fi \@xsect{#5}} \def\@ssect#1#2#3#4#5{\@tempskipa #3\relax \ifdim \@tempskipa>\z@ \begingroup #4\@hangfrom{\hskip #1}{\interlinepenalty \@M \sec@upcase{#5}\par}\endgroup \else \def\@svsechd{#4\hskip #1\relax #5}\fi \@xsect{#3}} \def\@startsection{paragraph}{4}{1em{\@startsection{paragraph}{4}{1em} {1ex plus .5ex minus .5ex}{-1em}{\bf}{\sec@upcase{Acknowledgments.}}} \let\acknowledgements=\@startsection{paragraph}{4}{1em \def\qanda@heading{Discussion} \newif\if@firstquestion \@firstquestiontrue \newenvironment{question}[1]{\if@firstquestion \section*{\qanda@heading}\global\@firstquestionfalse\fi \par\vskip 1ex \noindent{\it#1\/}:}{\par} \newenvironment{answer}[1]{\par\vskip 1ex \noindent{\it#1\/}:}{\par} \def\mathwithsecnums{ \@newctr{equation}[section] \def\hbox{\normalsize\arabic{section}-\arabic{equation}}}{\hbox{\normalsize\arabic{section}-\arabic{equation}}}} \def\section*{References{\section*{References} \bgroup\parindent=0pt\parskip=.5ex \def\relax{\par\hangindent=3em\hangafter=1}} \def\refpar\egroup{\relax\egroup} \def\section*{References{\section*{References} \list{\null}{\leftmargin 3em\labelwidth 0pt\labelsep 0pt\itemindent -3em \usecounter{enumi}} \def\relax{\relax} \def\hskip .11em plus .33em minus .07em{\hskip .11em plus .33em minus .07em} \sloppy\clubpenalty4000\widowpenalty4000 \sfcode`\.=1000\relax} \def\endlist{\endlist} \def\@biblabel#1{\relax} \def\@cite#1#2{#1\if@tempswa , #2\fi} \def\relax\refpar{\relax\relax} \def\@citex[#1]#2{\if@filesw\immediate\write\@auxout{\string\citation{#2}}\fi \def\@citea{}\@cite{\@for\@citeb:=#2\do {\@citea\def\@citea{,\penalty\@m\ }\@ifundefined {b@\@citeb}{\@warning {Citation `\@citeb' on page \thepage \space undefined}}% {\csname b@\@citeb\endcsname}}}{#1}} \let\jnl@style=\rm \def\ref@jnl#1{{\jnl@style#1\/}} \def\aj{\ref@jnl{AJ}} \def\araa{\ref@jnl{ARA\&A}} \def\apj{\ref@jnl{ApJ}} \def\apjl{\ref@jnl{ApJ}} \def\apjs{\ref@jnl{ApJS}} \def\ao{\ref@jnl{Appl.Optics}} \def\apss{\ref@jnl{Ap\&SS}} \def\aap{\ref@jnl{A\&A}} \def\aapr{\ref@jnl{A\&A~Rev.}} \def\aaps{\ref@jnl{A\&AS}} \def\azh{\ref@jnl{AZh}} \def\baas{\ref@jnl{BAAS}} \def\jrasc{\ref@jnl{JRASC}} \def\memras{\ref@jnl{MmRAS}} \def\mnras{\ref@jnl{MNRAS}} \def\pra{\ref@jnl{Phys.Rev.A}} \def\prb{\ref@jnl{Phys.Rev.B}} \def\prc{\ref@jnl{Phys.Rev.C}} \def\prd{\ref@jnl{Phys.Rev.D}} \def\prl{\ref@jnl{Phys.Rev.Lett}} \def\pasp{\ref@jnl{PASP}} \def\pasj{\ref@jnl{PASJ}} \def\qjras{\ref@jnl{QJRAS}} \def\skytel{\ref@jnl{S\&T}} \def\solphys{\ref@jnl{Solar~Phys.}} \def\sovast{\ref@jnl{Soviet~Ast.}} \def\ssr{\ref@jnl{Space~Sci.Rev.}} \def\zap{\ref@jnl{ZAp}} \let\astap=\aap \let\apjlett=\apjl \let\apjsupp=\apjs \def\hbox{$^\circ$}{\hbox{$^\circ$}} \def\hbox{$\odot$}{\hbox{$\odot$}} \def\hbox{$\oplus$}{\hbox{$\oplus$}} \def\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$<$}}}{\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$<$}}}} \def\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$>$}}}{\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$>$}}}} \def\hbox{\rlap{$\sqcap$}$\sqcup$}{\hbox{\rlap{$\sqcap$}$\sqcup$}} \def\hbox{$^\prime$}{\hbox{$^\prime$}} \def\hbox{$^{\prime\prime}$}{\hbox{$^{\prime\prime}$}} \def\hbox{$.\!\!^{\rm d}$}{\hbox{$.\!\!^{\rm d}$}} \def\hbox{$.\!\!^{\rm h}$}{\hbox{$.\!\!^{\rm h}$}} \def\hbox{$.\!\!^{\rm m}$}{\hbox{$.\!\!^{\rm m}$}} \def\hbox{$.\!\!^{\rm s}$}{\hbox{$.\!\!^{\rm s}$}} \def\hbox{$.\!\!^\circ$}{\hbox{$.\!\!^\circ$}} \def\hbox{$.\mkern-4mu^\prime$}{\hbox{$.\mkern-4mu^\prime$}} \def\hbox{$.\!\!^{\prime\prime}$}{\hbox{$.\!\!^{\prime\prime}$}} \def\hbox{$.\!\!^{\scriptscriptstyle\rm p}$}{\hbox{$.\!\!^{\scriptscriptstyle\rm p}$}} \def\hbox{$\mu$m}{\hbox{$\mu$m}} \def\onehalf{\hbox{$\,^1\!/_2$}} \def\hbox{$\,^1\!/_3$}{\hbox{$\,^1\!/_3$}} \def\hbox{$\,^2\!/_3$}{\hbox{$\,^2\!/_3$}} \def\hbox{$\,^1\!/_4$}{\hbox{$\,^1\!/_4$}} \def\hbox{$\,^3\!/_4$}{\hbox{$\,^3\!/_4$}} \def\ubvr{\hbox{$U\!BV\!R$}} \def\ub{\hbox{$U\!-\!B$}} \def\bv{\hbox{$B\!-\!V$}} \def\vr{\hbox{$V\!-\!R$}} \def\ur{\hbox{$U\!-\!R$}} \newcount\lecurrentfam \def\LaTeX{\lecurrentfam=\the\fam \leavevmode L\raise.42ex \hbox{$\fam\lecurrentfam\scriptstyle\kern-.3em A$}\kern-.15em\TeX} \def\plotone#1{\centering \leavevmode \epsfxsize=\textwidth \epsfbox{#1}} \def\plottwo#1#2{\centering \leavevmode \epsfxsize=.45\textwidth \epsfbox{#1} \hfil \epsfxsize=.45\textwidth \epsfbox{#2}} \def\plotfiddle#1#2#3#4#5#6#7{\centering \leavevmode \vbox to#2{\rule{0pt}{#2}} \special{psfile=#1 voffset=#7 hoffset=#6 vscale=#5 hscale=#4 angle=#3}} \newif\if@finalstyle \@finalstylefalse \if@finalstyle \ps@myheadings \let\ps@title=\ps@paspcstitle \else \ps@plain \let\ps@title=\ps@plain \fi \ds@twoside \makeatother \begin{document} \title{Stellar Orbits in Barred Galaxies with Nuclear Rings} \author{Clayton H. Heller} \affil{Universit\"ats Sternwarte, Geismarlandstra\ss e 11, D-37083 G\"ottingen, Germany} \author{Isaac Shlosman} \affil{Department of Physics and Astronomy, University of Kentucky, Lexington, KY 40506-0055, USA} \begin{abstract} We investigate the dynamical response of stellar orbits in a rotating barred galaxy potential to the perturbation by a nuclear gaseous ring. The change in 3D periodic orbit families is examined as the gas accumulates near the inner Lindblad resonance. It is found that the $x_2/x_3$ loop extends to higher Jacobi energy and a vertical instability strip forms in each family. These strips are connected by a symmetric/anti-symmetric pair of $2\!\!:\!\!2\!\!:\!\!1$ 3D orbital families. A significant distortion of the $x_1$ orbits is observed in the vicinity of the ring, which leads to the intersection between orbits over a large range of the Jacobi integral. We also find that a moderately elliptical ring oblique to the stellar bar produces significant phase shifts in the $x_1$ orbital response. \end{abstract} About 2/3 of all disk galaxies are weakly or strongly barred (de Vaucouleurs 1963), many more are ovally distorted (Bosma 1981; Kormendy 1982) or have triaxial bulges (Kormendy 1994). Central starburst activity in these galaxies often delineates $\sim$few$\times100$ pc size ring-like structures of star forming regions mixed with molecular gas and dust (Buta \& Crocker 1993). Nuclear ``rings'' seem to be associated with inner Lindblad resonances (ILRs) (Telesco \& Decher 1988; Shlosman\,{\it et\,al.}\,1989; Kenney\,{\it et\,al.}\,1992; Athanassoula 1992; Knapen\,{\it et\,al.}\,1995a,b). Their intrinsic shapes vary from circular to moderately elliptical, in which case they lead stellar bars by $\sim 50-90\hbox{$^\circ$}.$ As such, nuclear rings are moderately strong perturbations on the gravitational potential of the central galactic region, thus affecting stellar orbits and gas flow there. We analyze the main stellar orbits (in the plane and 3D) in the presence of a ring (see also Heller \& Shlosman 1995). The galaxy model consists of the superposition of four components: disk, bulge, bar, and ring. The disk is represented by a Miyamoto-Nagai potential, the bulge by a Plummer sphere, and the bar by a triaxial Ferrers density distribution. The ring is centered in the ILR region and for the models presented here is equivalent to $\sim10^9\,{\rm M}_\odot$ or 38\% of the local mass (Model D). The characteristic diagram for the main planar prograde periodic orbits in the inner region of the model when no ring is present (Model A) is shown in the upper frame of Figure~1. \begin{figure}[t] \plotfiddle{ringfig1.ps}{5cm}{0}{80}{55}{-240}{-200} \caption{Characteristic diagrams of the $x_1$, $x_2$, and $x_3$ families for models without (A) and with (D) a nuclear ring. Stable sections of the characteristics are represented by solid lines while unstable are broken. Four vertical instability strips are marked. The long-dashed curve is the zero velocity curve.} \end{figure} The three direct families $x_1$, $x_2$, and $x_3$ are shown along with two vertical instability strips. From the vertical instability strips bifurcate pairs of symmetric/anti-symmetric 3D orbital families, $2\!\!:\!\!2\!\!:\!\!1$ (BAN/ABAN) and $2\!\!:\!\!3\!\!:\!\!1$ families from $S_1$ and $S_2$, respectively. The $2\!\!:\!\!3\!\!:\!\!1$ families have interesting orbital shapes that are symmetric about one vertical plane while being anti-symmetric about the corresponding perpendicular vertical plane. As the mass of the ring is increased a ``bump'' in the $x_1$ family forms and broadens at an ${\rm E}_{\rm J}$ below the ILR. This distortion represents a local maximum in the y-extent of the orbits, resulting in a large region of the $x_1$ family to have orbits that intersect with other $x_1$ orbits at higher ${\rm E}_{\rm J}$. Such orbit intersections also occur in the $x_2$ family as a local maximum in orbit eccentricity develops along the sequence. Also, as the ring's mass is increased the region of stability close to the plane of the $2\!\!:\!\!2\!\!:\!\!1$ symmetric family increases while the $x_2/x_3$ loop extends to higher ${\rm E}_{\rm J}$ and develops two regions of vertical instability. These two instability strips, one on $x_2$ and one on $x_3$, are connected by a symmetric/anti-symmetric pair of $2\!\!:\!\!2\!\!:\!\!1$ families elongated perpendicular to the stellar bar. The symmetric family is stable over half of its characteristic, while the anti-symmetric is unstable everywhere. The bottom frame of Figure~1 shows the planar characteristic diagram for the model with a circular ring and indicates the location of the $x_2/x_3$ instability strips $S_3$ and $S_4$. In Figure~2a \begin{figure}[t] \plotfiddle{ringfig2a.ps}{2cm}{0}{33}{32}{-210}{-115} \plotfiddle{ringfig2b.ps}{2cm}{0}{40}{33.1}{-35}{-87} \caption{(a) Twisting of $x_1$ orbits in model with oblique elliptical ring. The ring with an ellipticity of 0.4 and semi-major axis 0.04 units is leading the bar by 60\,deg. The frame is 2\,Kpc on a side. (b) Eccentricity and position angle of $x_1$ orbits from models without ring (dot-dashed), with circular ring (dashed), and oblique elliptical ring (solid), as a function of the Jacobi energy.} \end{figure} we show the phase shift or twisting of the $x_1$ orbital alignment in response to a moderately elliptical ($e=0.4$) ring leading the stellar bar by 60 degrees. The change in ellipticity and position angle is given in Figure~2b and is compared with models A and D. It can be seen that while the eccentricity as a function of ${\rm E}_{\rm J}$ is only slightly offset from the circular ring case, the position angle of the orbit semi-major axis swings from -10 to 35 degrees with respect to the bar. Note, that innermost $x_1$ orbits trail the bar. The interior orbits remain stable and continue to trap a significant region of phase space around them. The main effect of the circular nuclear ring is to produce intersecting orbits over a wide range of Jacobi energies in both the $x_1$ and $x_2$ orbit families. Gas on such orbits will shock and dissipate energy on a dynamical time scale. As a consequence, the gas will quickly settle down deep inside the resonance region, further enhancing the ring. The growth of the ring is limited by its self-gravity. For a non-circular ring oblique to the stellar bar and leading it, the twisting of the $x_1$ orbits will further enhance shocks in the gas. It is clear from Figure 2, that both trailing and leading shocks will develop.
\section{Introduction} Cosmic strings may have formed at a phase transition in the early universe \cite{Tom,Review}. Information about the initial statistics of a string network, after the point at which thermal fluctuations become unimportant and the strings are `frozen in', has largely emerged from the numerical simulations first performed by Vachaspati and Vilenkin \cite{VV}. The simplest case, involving the spontaneous breaking of a $U(1)$ symmetry, is mimicked by assigning a phase between 0 and $2\pi$ to each point on a regular lattice. The lattice spacing then corresponds to a correlation length $\xi$ characteristic of the scalar field acquiring the non-zero vacuum expectation value. To look for field configurations with non-trivial topology, the `geodesic rule' is invoked. This proposes that to minimise gradient energy the field will follow geodesic paths on the vacuum manifold as a path in configuration space is traversed. The phase will thus follow the `shortest path' between values on adjacent lattice sites. A winding of $\pm 2\pi$ around a plaquette in the lattice means that a line-like distribution of zeroes of the field will pierce it --- a cosmic string. If we impose periodic boundary conditions on our lattice, it becomes obvious that all string must be in the form of closed loops. One would expect that this procedure gives a string configuration in our box that is statistically similar to that when neighbouring, causally disconnected regions are present\cite{Ray}. {}From the distribution of lengths of loops, it is easy to make a distinction between `infinite' string (winding around the box many times) and smaller loops, peaked at the minimum size of four lattice spacings. The analytic form of the small-loop distribution is well understood statistically. It is found that, for a cubic lattice, around $70-80\%$ of string exists as infinite string in this scenario, which has long been used as the generator of initial configurations for the numerical evolution of string networks \cite{AS,AT,BB}. If a non-minimal discretisation of the vacuum manifold is used (that is, in the case of a broken $U(1)$ symmetry, approximating $S^1$ with more than the smallest number of points, $\theta =$ 0, $2\pi /3$, $4\pi /3$) and we employ a cubic lattice of points, in principle it is possible for all six faces of a fundamental cell to contain strings. Even in the minimal case, it is possible for four faces to do so. This requires a random choice to be made, pairing the incoming and outgoing strings. The only method that avoids this ambiguity is to use a tetrahedral lattice with a minimal discretisation, so that at most one string enters and leaves each cell. String configurations arising from this model have been analysed recently \cite{HS} and it is found that a slightly lower fraction (around $65\%$) exists as infinite string. As a string network evolves, by means of intercommutation and expansion of the universe, it has been predicted and (to different extents) observed in the simulations that the characteristic lengths describing it approach a `scaling regime', in which they grow in proportion to the horizon size. A typical evolving network will display an initial flurry of loop production before settling into this scaling regime with a few long strings and large loops per horizon volume continuing to (self-) intersect and produce smaller loops. As such, it has been supposed that the initial details of the string network are largely washed out after a few expansion times. This indeed seems to be the case, from both the numerical work and more recent analytic models of network evolution \cite{BigTom}. The question arises, though --- is there a causal mechanism for creating a string distribution with significantly less infinite string? In the most extreme case, it is possible that if there were no infinite string at all, all the loops would disappear within a finite (and quite short) time. Recent work by Ferreira and Turok \cite{Ped} partly confirms this, showing that a different type of scaling occurs in that case. One way of testing this idea is to attempt to rectify one of the major simplifications inherent in the Vachaspati-Vilenkin algorithm, and introduce a {\it distribution} of domain volumes in the initial conditions, instead of simply assuming that causally disconnected regions of one value of the field are of equal volume ($\sim \xi^3$). \section{Implementing the algorithm} A cubic lattice was used with a near-continuous \footnote{ i.e. a very high-density discretisation of the circle, that allows the use of integer arithmetic} representation of the vacuum manifold, despite the reduction in ambiguity that can be achieved with the tetrahedral lattice, as mentioned above, since it simplified the process of creating a domain structure. Physical space is partitioned into regions of constant $U(1)$ phase by throwing down domains of random diameter within specified limits and gradually covering the lattice, dealing with the overlap and fragmentation of these regions as the box becomes filled with the broken phase. Roughly spherical domains were experimented with at first. However, after taking account of the significant domain overlap that resulted from the random filling of the lattice, and to make the task of ensuring that no domain was created entirely within another more straightforward, cubical domains were used. Once this was completed, strings were located and traced through the lattice, following the edges of either three or four adjacent domains. It should be made clear that this is no more than a means of setting up a domain structure, and in no way claims to simulate the dynamics of an actual phase transition. Indeed, it is not obvious what order of transition the results of this algorithm apply to, though it would seem more closely related to string formation at the interfaces of expanding bubbles of the true vacuum, rather than the uniform emergence of a domain structure in a second-order transition. As a first guess we might expect a Gaussian distribution of domain volumes, peaked around some mean value. As it happens this is difficult to realise, and the size distribution appears to be more Poissonian (Fig.\ 1). The results are interesting nevertheless, and in particular the fact that the resulting form of the graph seems largely insensitive to modifications to the domain-laying algorithm. The range of sizes of domains laid down was systematically increased in order to plot the fraction of the total string density as `infinite' string, $f_{\infty}$, against domain volume variance. The variance is normalised to the mean domain volume in order to remove effects due to uniform scaling-up of domain volumes. In the zero-variance limit the Vachaspati-Vilenkin result of $f_{\infty}\simeq 0.76$ is obtained. The precise value is weakly dependent on the imposed loop/`infinite' string cutoff --- if we set the maximum size of a loop to be $4N$, $10N$ and $N^2 /2$, where $N$ is the length of the side of our box in units of the smallest possible domain size, the zero-variance values of $f_{\infty}$ are 0.78, 0.77 and 0.75 respectively. Higher-variance values of $f_{\infty}$ change by a similar amount. A cutoff of $N^2 /4$ was used in the plots presented here. \begin{figure*} \label{fig:hist} \begin{minipage}{4.5in} \setlength{\unitlength}{1in} \begin{picture}(2,4.5) \put(0.5,0.40){\psfig{file=fig1b.ps,width=2in}} \put(1.5,0.20){(b)} \put(0.5,2.6){\psfig{file=fig1a.ps,width=2in}} \put(1.5,2.4){(a)} \end{picture} \end{minipage} \caption{Histograms of domain volumes. Each illustrates range of domain sizes present for a particular realisation of the laying algorithm, which fills the box with domains of diameter randomly chosen in the range 1 to $D$. Figure (a) shows the results for a $100^3$ box with $D$=5; (b) $D$=15.} \end{figure*} \section{Results} In figure \ref{fig:main}, each point is the average of 20 runs, with fixed limits on the range of sizes of domains laid down. The first point is the result of filling half the box with domains of side 1 or 2, randomly chosen. The remaining space is filled with unit domains, in order to achieve a low volume variance. Subsequent points correspond to the box being filled entirely with domains of sides between 1 and $D$ ($D$=2,3,...,18). \begin{figure} \centerline{\psfig{file=fig2.ps,width=3.2in}} \caption{Results for $100^3$ lattice.} \label{fig:main} \end{figure} The initial decrease in $f_{\infty}$ with increasing variance is perhaps intuitively understood. Typically, on a regular lattice, the string performs a self-avoiding random walk - one in which the string is not permitted to intersect itself or any other strings except at the `origin'. This property is imposed on the strings by the restrictions of the lattice method. Such a walk has the property that the end-to-end distance $l$, in units of the step length, is related to the mean displacement $R$ by \[ R \sim l^{3/5}. \] In fact, the presence of other strings provides an extra repulsive effect and so gives the string near-Brownian characteristics $(R \sim l^{1/2})$. This is confirmed in the simulations --- typical figures for the $l$ exponent were $\simeq 0.47 \pm 0.04$ at zero variance. However, the presence of extended regions of space from which the string is excluded, i.e. larger domains, provides restrictions on the ability of the string to `fold in' on itself. Effectively, in the region of these larger domains, we expect that a loop of a given radius will have a smaller perimeter than a similar loop in a region of unit domains. This will increase the density of loops below the cutoff size. It is worth noting that as the variance increases, there exist more ways to fill the box and so a wider range of possible domain configurations. This also emphasises the point that volume variance is almost certainly not the only parameter describing the spatial distribution of phases that determines $f_{\infty}$. Statistics at high values of $N$ became unreliable, but it is intriguing to speculate whether further increases in the variance could well force all string to be in the form of small loops. The finite size of the simulation limits the maximum value of N we can reasonably investigate. \section{Percolation effects} Another way to observe a reduction in the density of infinite string is to impose a `tilt' on the vacuum manifold, statistically favouring the occurrence of one phase \cite{Tan}. \begin{figure} \centerline{\psfig{file=fig3.ps,width=2.8in}} \caption{Plot of $f_{\infty}$ against the bias parameter $\gamma$ for a zero-variance, $100^3$ box. Each point is averaged over 20 realisations.} \label{fig:perc} \end{figure} If we employ a three-point discretisation and gradually increase a bias parameter $\gamma$, such that $ P({\mbox{phase 1}}) = \gamma, P(2)=P(3)= \frac{1}{2}(1-\gamma) $, we find that $f_{\infty}$ drops smoothly, reaching zero at $\gamma \simeq 0.5$ (Fig.\ \ref{fig:perc}). A phase is said to percolate when it is possible to trace continuous `infinite' paths through that phase in the box. Clearly there {\it is} a relation between the strings percolating (passing through the Hagedorn transition) and the percolation of phases in the box in this minimally discretised case. For a string to exist it must have all three phases around it. An infinite string will therefore ensure that all phases percolate (including diagonally adjacent regions). At least one phase percolates for all values of $\gamma$ --- the critical probability for the occurrence of one phase $p_c$, above which it percolates, is 0.31 \cite{VV}. In fact, we note that when $\gamma \simeq 0.5$, $P(2) = P(3) \simeq 0.25$, which is very near the percolation threshold. It is interesting to ask whether there is a connection between increasing the variance of the domain volume and moving away from string percolation. Statistical fluctuations in the volume of the domains will result in the fractions of box volume occupied by each phase departing from $1/3$, becoming more divergent as the range of sizes of domains increases. We suggest that this can be interpreted as an effective tilt of the vacuum manifold. The fact that a bias will reduce the amount of long string is easily understood. We consider the probability $p$ that a given plaquette is pierced by a string ($p=8/27$ ($\approx 0.30$) in the case of three-point discretisation). Given that we have an ingoing string through one face, what is the probability that this string will turn through $\pi /2$ in the cell under consideration? Obviously the opposite face has four independent phases (1,2 or 3) attached to it, so the probability of it containing an {\it outgoing} string is $p/2$. Given this configuration, the probability of the cell containing a further ingoing/outgoing string is 1/4. Thus the probability that our string continues through the cell undeviated, assuming we pair strings within the box randomly, is \[ \left( \frac{p}{2} \times \frac{3}{4} \right) + \left( \frac{p}{2} \times \frac{1}{4} \right) \times \frac{1}{2} = \frac{7}{16}p. \] As we increase the bias parameter $\gamma$ we obviously decrease the probability $p$. In fact, \[ p = 2 \gamma (1- \gamma)^2,\] giving values for $p$ of 0.30, 0.15 and 0.02 for $\gamma = 1/3$, 2/3 and 8/9 respectively. Thus, strings will be more likely to fold up as $\gamma$ grows, and the population of small loops will increase. This agrees with ref. \cite{HS}, who point out that as the bias increases and the strings stop percolating, the fractal dimension of the strings becomes higher than two and they tend to `crumple up' more --- they become self-seeking random walks. Unfortunately, statistics were too poor to investigate any change in the fractal dimension of the strings as the bias or the variance was increased. \begin{figure} \centerline{\psfig{file=fig4.ps,width=2.8in}} \caption{Correlating the volume variance with the bias parameter $\gamma$} \label{fig:correlate} \end{figure} However, it is possible to investigate qualitatively the connection between domain volume variance and a vacuum tilt by calculating values for the variance for each value of the bias $\gamma$ in figure \ref{fig:perc}. We set up the box by throwing down phases in single cells according to the biased probability distribution. We then group adjoining cells containing the same phase to form larger domains, whose volumes we calculate. The results are plotted in figure \ref{fig:correlate}. The errorbars are misleading since there is clearly a correlation, and this is to be expected intuitively --- the more one phase appears at the expense of others, the bigger the range of sizes of domains present. Exploring the idea further, we calculate the bias parameter `geometrically', given the volume occupied by each of the three phases in the box. The results of this procedure are shown in figure \ref{fig:effective_tilt}. The values of $\gamma_{\mbox{eff}}$ are too low to correspond to those in the original figure (\ref{fig:perc}), and the plots are not similar in form. However, the increase of $\gamma_{\mbox{eff}}$ as $f_\infty$ decreases is in agreement. \begin{figure} \centerline{\psfig{file=fig5.ps,width=2.8in}} \caption{Calculation of the effective bias parameter in the minimally-discretised case. For increasing values of N, a value for $\gamma$ was calculated from the fractions of the box occupied by each of the three phases.} \label{fig:effective_tilt} \end{figure} \section{Conclusions} We have seen that a simple extension to the accepted numerical model for string formation can yield a significantly different estimate of the amount of infinite string present. It seems feasible that increasing the variance of the volumes of regions with different VEVs is equivalent to an effective tilt of the vacuum manifold, that leads to a reduction in the density of infinite string when considering a finite volume with periodic boundary conditions. Whether this is the case in the infinite-volume limit is more debatable. With three-point discretisation, the variance becomes ill-defined in this regime, as all three phases percolate. However, in this limit it is also unclear whether there is truly a population of infinite string, distinct from the $l^{-5/2}$ loop distribution, or if it is purely an artefact of the boundary conditions. As yet there is no physical argument to suggest what the volume variance in a given phase transition will be --- and even, considering the effects of phase equilibration at domain boundaries, how well-defined this quantity is. Models of dynamic defect formation, even with simplified treatments of the physics involved in a real phase transition, may give improved predictions of the defect configurations \cite{Julian}. One of the consequences of the existence of GUT-scale strings is the possibility of their being responsible for structure formation. It is only the infinite string and large loops that will survive long enough to be useful in this scenario, as a huge number of Hubble times elapse between string formation and when perturbations on interesting (galactic) scales will begin to grow. It may be that if the amount of long string present is very low, then their structure-seeding properties will be less significant than previously thought. Certainly, the proposed existence of a unique scaling solution for the string network, independent of initial conditions, would be put into doubt. \section*{Acknowledgements} The authors would like to thank Pedro Ferreira and Julian Borrill for helpful discussions, and James Robinson for contributing part of the code. A.Y. was funded by PPARC.
\section{Introduction} \label{sec:intro} The symmetry of the order parameter $\Delta_{\bf q}$---the energy gap in the quasiparticle excitation spec\-trum---in the high-$T_c$ superconductors (HTSC's) has been a very active area of research in the last few years. Assuming that the order parameter is a spin singlet as seems to be indicated by Knight-shift experiments \cite{Barrett}, the angular pairing state has to be even, leaving $s$-wave ($L$=0) and $d$-wave ($L$=2) as the main alternatives. Different theories make different predictions for the order parameter, which is the reason for the interest in the symmetry of $\Delta_{\bf q}$. There are essentially two classes of experiments to determine the symmetry of the energy gap, those that are sensitive to the magnitude of the order parameter, and those that are sensitive to the phase of it. Measurements of the acoustic attenuation \cite{MorseBohm} and the NMR nuclear-spin relaxation rate \cite{Hebel}, which belong to the first category, were instrumental in establishing the BCS theory in the conventional superconductors \cite{Schrieffer}. Acoustic Attenuation was also used to examine anisotropy in the conventional superconductors due to the crystal symmetry \cite{Morse}. However, while many NMR experiments have been performed on the HTSC's (cf. the review by Slichter {\em et al.}\ \cite{Slichter93}), acoustic attenuation experiments have been neglected in recent years. We will discuss the theory of acoustic attenuation for the HTSC's, and propose an experiment to probe the symmetry of the order parameter using acoustic attenuation. Acoustic attenuation is observed with an experimental setup \cite{Rayne}, where an ultrasonic signal, typically with frequencies ranging from 100 MHz to 10 GHz, is fed into a sample through a transducer quartz. The resulting phonons with wave vector $\bf q$, energy $\omega$, and polarization $\lambda$ can be scattered by quasiparticles, and the remaining fraction of phonons is observed at the other end of the sample. In addition to the electronic mechanism, where the phonons are scattered by quasiparticles, there will also be a lattice contribution to the attenuation rate. We will discuss how the electronic contribution can be resolved against the lattice background. \section{Acoustic Attenuation} \label{sec:acat} We consider processes in which phonons are absorbed by a sample through scattering of quasiparticles from a state ${\bf p_1}$ into a state ${\bf p_2}$, where ${\bf q} = {\bf p_2} - {\bf p_1} + {\bf K}$ is the momentum of the incoming phonon and ${\bf K}$ is a reciprocal lattice vector. The inverse process of spontaneous emission of phonons by quasiparticles also has to be taken into account. To understand the effect of anisotropy of the order parameter it is important to understand how the involved momenta are related. The quasiparticle is scattered from a state $\bf p_1$ near the Fermi surface to another state $\bf p_2$, which also has to be near the Fermi surface. This means that both ${\bf p_1}$ and ${\bf p_2}$ are of order $k_F$. $\bf q$, the phonon momentum, is smaller by some orders of magnitude. This means that $\bf p_1$ and $\bf p_2$ point essentially in the same direction. The HTSC's are quasi-two-dimensional layered materials. Their Fermi surface is nearly cylindrical with little dependence on the coordinate in the $c$-direction. For a three-dimensional Fermi sphere, $\bf p_1$ and $\bf p_2$ are fixed on a belt around the Fermi surface \cite{Morse}, perpendicular to the direction of the phonon momentum $\bf q$. For a given phonon momentum $\bf q$ in the $a$-$b$-plane in the quasi-two-dimensional case, this belt degenerates to two points on opposite sides of the Fermi surface. Thus, the order parameter is probed only in a specified direction. This provides an opportunity to measure the anisotropy in the magnitude of the order parameter. The interaction of phonons with quasiparticles can be written as \begin{equation} {\cal H}_{\mbox{el-ph}} = \sum_{{\bf p_1},{\bf p_2},s,\lambda} g_{{\bf p_1,p_2},\lambda} (a_{{\bf q} \lambda} + a_{-{\bf q} \lambda}^\dagger) c_{{\bf p_2} s}^\dagger c_{{\bf p_1} s}, \label{eq:HamiltonianAc} \end {equation} where the relation between $\bf q$, $\bf p_1$, and $\bf p_2$ was given above. Here $g_{{\bf p_1,p_2},\lambda}$ is the interaction strength, $a_{{\bf q} \lambda}$ is a phonon destruction operator, and $c_{{\bf p_1} s}$ an electron destruction operator. The electron operators $c$, $c^\dagger$ can be transformed into superconducting quasiparticle operators $\gamma$, $\gamma^\dagger$ with a standard Boguljubov transformation. Starting from this Hamiltonian it is straightforward to derive an expression for the acoustic attenuation rate (e.g.\ cf.\ Schrieffer \cite{Schrieffer}) $\alpha_{{\bf q},\lambda}$ with \begin {equation} \alpha_{{\bf q},\lambda}=4\pi \sum_{{\bf p_1},{\bf p_2}} |g_{{\bf p_1,p_2},\lambda}|^2 \, n^2 ({\bf p_1},{\bf p_2}) \, (f_{\bf p_1}-f_{\bf p_2}) \, \delta(E_{\bf p_2}-E_{\bf p_1}-\omega_{{\bf q} \lambda}) \end{equation} for phonons of wave vector $\bf q$ and polarization $\lambda$. The involved phonon frequency $\omega_{{\bf q} \lambda}$ is given by the dispersion relation of the phonons. The square of the so-called coherence factor $n$ is evaluated to be \begin {equation} n^2({\bf p_1},{\bf p_2})=(u_{\bf p_1} u_{\bf p_2} - v_{\bf p_1} v_{\bf p_2})^2= \frac {1} {2} \left(1+ \frac {\epsilon_{\bf p_1} \epsilon_{\bf p_2}-\Delta_{\bf p_1} \Delta_{\bf p_2}} {E_{\bf p_1} E_{\bf p_2}}\right). \end{equation} Here, ${\epsilon_{\bf p}}$ is the single-particle energy relative to the Fermi energy and $E_{\bf p}$ is the superconducting quasiparticle excitation energy, $E_{\bf p}=\sqrt{\epsilon_{\bf p}^2+ \Delta_{\bf p}^2}$. $f_{\bf q} \equiv 1/(1+\exp (E_{\bf q}/k_B T))$ is the Fermi function. We assume that $g_{{\bf p_1,p_2} ,\lambda}$ only depends on the momentum transfer ${\bf q} = {\bf p_2} - {\bf p_1}$. It is assumed that the order parameter depends only on the direction, but not on the magnitude of {\bf p}, and that the angular dependence of the order parameter does not change with temperature $T$. Finally we are led to \begin {eqnarray} \alpha_{{\bf q},\lambda} & = & \mbox{const.} \times \int d\epsilon_{\bf p_1} d\epsilon_{\bf p_2} \left(1-\frac {\Delta_{\bf p_1} \Delta_{\bf p_2}} {E_{\bf p_1} E_{\bf p_2}}\right) \nonumber \\* & & \times (f(E_{\bf p_1})-f(E_{\bf p_2})) \delta(E_{\bf p_2}-E_{\bf p_1}-\omega_{q \lambda}) \nonumber \\* & = & \mbox{const.} \times \int dE_{\bf p_1} \frac {E_{\bf p_1}} {\sqrt{E_{\bf p_1}^2 -\Delta_{\bf p_1}^2}} \frac {E_{\bf p_1}+\omega_{q \lambda}} {\sqrt{(E_{\bf p_1}+\omega_{q \lambda})^2-\Delta_{\bf p_2}^2}} \nonumber \\* & & \times \left(1-\frac {\Delta_{\bf p_1} \Delta_{\bf p_2}} {E_{\bf p_1} (E_{\bf p_1}+\omega_{q \lambda})}\right) (f(E_{\bf p_1})-f(E_{\bf p_1}+\omega_{q \lambda})). \label{eq:int} \end{eqnarray} We do not get an angular integration because $\bf q$ picks certain values for the directions of $\bf p_1$ and $\bf p_2$. These directions, $\theta_1$ and $\theta_2$, are very close to one another due to momentum conservation as argued above, and can be controlled experimentally. For our calculations we used $\hbar \omega_{q \lambda} = 10^{-5}k_B T_c$. The results are quite insensitive to the value of $\omega$ as long as $\hbar \omega << k_B T_c$. \section{Symmetry of the Energy Gap} \label{sec:symmetry} To calculate the acoustic attenuation rate $\alpha_{{\bf q},\lambda}(T)$, the temperature dependence of the reduced gap, $\Delta_0(T)/\Delta_0(0)$, will be assumed to be BCS-like as a function of reduced temperature. We have tested that this assumption is consistent with both an $s$- and $d$-wave gap with appropriate potentials in the BCS gap equation \cite{Diplom}. For the angular dependence of the gap, $\Delta({\theta})$, different models will be examined: \begin {equation} \Delta(\theta,T) = \left\{ \begin{array} {ll} \Delta_0(T) &\mbox {isotropic } s\mbox{-wave},\\ \Delta_0(T) \cos (2\theta) &d\mbox{-wave},\\ \Delta_0(T) \left[ a\cos^2 (2\theta)+(1-a) \right] &\mbox {anisotropic } s\mbox{-wave}. \end{array} \right. \label{eq:gapforms} \end {equation} In conventional superconductors one calculates the attenuation rate in the superconducting state normalized to the attenuation rate in the normal state. In HTSC's, however, this is not an interesting quantity because it is not experimentally accessible. In conventional superconductors one can always drive the system into the normal state even at temperatures much below $T_c$ by applying a sufficiently large magnetic field. In HTSC's, however, the critical fields are prohibitively high. Thus, we normalize the attenuation rate to its value at $T_c$. Evaluating Eq.\ (\ref{eq:int}) with $\bf q$ pointing in different directions relative to the lattice amounts to taking different effective magnitudes of $\Delta$. We consider $q/k_F$ very small so that $\theta_1\simeq\theta_2$, and thus $\Delta_{\bf p_1}\simeq\Delta_{\bf p_2}$. The direction of $\bf q$ can be controlled experimentally. Fig.\ 1 shows the attenuation rate as a function of temperature for different effective magnitudes of the gap. A $2 \Delta_0(T=0)/k_B T_c=3.5$ corresponds to the isotropic BCS case. Because for a particular $\bf q$ only the magnitude of the gap in the direction perpendicular to $\bf q$ enters, Fig.\ 1 holds for a particular $\bf q$ no matter what the symmetry of the gap is. This is true as long as the temperature dependence of the gap in that direction behaves like the BCS temperature dependence. At the high temperatures near $T_c$, $T_c \simeq 90$ K, lattice contributions to the attenuation rate become important, but the strong temperature dependence of the electronic contribution near $T_c$ provides an opportunity to still measure it. Since the superconducting transition should not affect any but the electronic contribution to the attenuation, measuring the attenuation rate just above and below $T_c$ allows one to separate {\em electronic} contributions from {\em lattice} contributions. In particular any eventual anisotropy of a lattice contribution should not be influenced by the superconducting transition, at least as long as the temperature difference between the measurements is not too large. Using this approach, the anisotropy of the electronic attenuation rate becomes an experimentally accessible quantity, which can be used to examine the symmetry of the superconducting gap in the HTSC's. For an isotropic $s$-wave gap the electronic contribution to the attenuation rate should not change as the crystal is rotated. For an anisotropic order parameter, either $s$- or $d$-wave, maxima should be observed in directions perpendicular to where the order parameter is a minimum or even has nodes. At a node the attenuation rate should go up to the normal state value at the corresponding temperature. Fig.\ 2 shows how the acoustic attenuation rate varies at a temperature of 0.95 $T_c$ for different symmetries of the order parameter as a function of the direction of the incoming phonons relative to the lattice. A value of $2\Delta_0(T=0)/k_B T_c=6$ is assumed. All rates are normalized to the acoustic attenuation rate without a gap, which is essentially the electronic attenuation rate at $T_c$, since normal electronic contributions should not vary much over this small range of temperatures. While all symmetries show a significant suppression in certain directions, which allow the resolution of the electronic contribution against the lattice background, the very anisotropic symmetries show little or no suppression in those directions where minima or nodes of the gap are located. \section{Discussion and Summary} \label{sec:disc} The acoustic attenuation method, which was very successful in verifying BCS theory for conventional superconductors, has the potential to provide useful information on the order parameter in high $T_c$ materials. Early on some measurements have been made (Yusheng {\em et al.}\ \cite {Yusheng}), but it has been argued \cite {Almond} that the effect seen was too large to be the actual electronic contribution to the attenuation rate. However, since then sample qualities have been improved considerably, and, with our proposed focusing on temperatures near $T_c$, this method has a potential that has not been exploited yet. Won and Maki \cite {Won} recently also discussed acoustic attenuation in HTSC's; however, they do make some additional approximations to solve the problem analytically, only consider $d$-wave, and argue that at {\em low} temperatures, where the rate is already strongly suppressed, this rate should be strongly anisotropic. We propose that the attenuation be measured at temperatures near $T_c$, where one should see the anisotropy, but still have a measurable rate. Any observed sharp drop in the attenuation rate near $T_c$ can be attributed to the electronic properties of the system since the lattice properties should not change dramatically near the superconducting transition. {\bf Note Added in Proof:} After the present paper was submitted for publication, we learned of a paper by Kostur {\it et al.} \cite{Kostur} (KBF) which reports on calculations of ultrasonic attenuation in a model d-wave superconductor. The results are quite similar to those presented here. We feel that our paper compliments KBF in that we consider anisotropic s-wave superconductors in addition to d-wave. Also, we suggest that the best way to separate the anisotropic attenuation due to quasiparticles from that due to the lattice is to compare results just above $T_c$ with those just below $T_c$, whereas KBF seem to advocate looking at low temperatures. \section*{Acknowledgements} We would like to thank J.\ Appel and C.\ Timm for useful discussions. T.\ W.\ likes to thank Indiana University for the hospitality during his stay in Bloomington, and gratefully acknowledges financial support through the Indiana University Overseas Exchange Fellowship. \pagebreak
\section{Introduction} In a preceding article \cite{p10}, an attempt is made to simplify the study of the algebraic structure of dynamical systems involving ($p=2$) parabose and parafermi variables. The approach presented in \cite{p10} is aimed to facilitate the analysis of systems possessing parabose -- parafermi supersymmetry, thus providing the necessary framework for investigating the relation between the conventional parastatistics of Green \cite{green} and the more recent developments of parasupersymmetric quantum mechanics \cite{r-s,b-d,p8,p9}. More specifically the purpose of the present article is to answer to the question: \begin{itemize} \item[] {\em Is parabose -- parafermi supersymmetry the same as parasupersymmetry?} \end{itemize} One should note that the so-called ``parasupersymmetric oscillators'' studied in the literature, e.g., \cite{r-s,b-d}, are constructed using some specific matrix representation of the parafermi operators. The analysis presented in this paper does not restrict to matrix representations and treats the parafermi and parabose operators (variables) as fundamental mathematical objects. The paper is organized as follows: In Sec.~2, the main results of \cite{p10} are quoted and the possibility of the existence of parabose -- parafermi (supersymmetry) transformations is investigated. In Sec.~3, the analogy between the ($p=2$) parabose -- parafermi supersymmetry and the ordinary bose -- fermi supersymmetry is discussed. The example of the supersymmetric oscillator is then reviewed and the ($p=2$) parabose -- parafermi oscillator is introduced by analogy. In Sec.~4, the parabose -- parafermi supersymmetries of the oscillator are studied. In Sec.~5, the super Lie algebra of the symmetries of the oscillator is used to define the notion of {\em ($p=2$) Supersymmetric Paraquantum Mechanics}. This section also offers a detailed treatment of the degeneracy structure of general ($p=2$) supersymmetric paraquantum mechanics. Sec.~6 is devoted to an analysis of the energy eigenstates and the spectrum degeneracy of the parabose -- parafermi oscillator. Here, the explicit form of a complete set of energy eigenstate vectors is obtained. Sec.~7 includes the conclusions. For brevity we shall use the notation $\pi b$, $\pi f$, $\pi SUSY$ for {\em ($p=2$) parabose, parafermi,} and {\em parabose -- parafermi supersymmetry}, and abbreviations SQM, PSQM, SPQM for {\em supersymmetric quantum mechanics}, {\em parasupersymmetric quantum mechanics}, and {\em supersymmetric paraquantum mechanics}, respectively. We shall follow Einstein summation convention of summing over repeated indices throughout the paper, unless otherwise indicated. \section{Algebraic Structure of Classical $\pi SUSY$} In this section, first we recall the constructions developed in \cite{p10}. The algebra of the creation $a_k^{\mu\dagger}$ and annihilation operators $a_k^{\mu}$ for the ($p=2$) $\pi b$ ($\mu=0$) and $\pi f$ ($\mu=1$) variables is given by: \begin{eqnarray} a_k^{\mu}&=&\sum_{\alpha=0}^1\zeta^{\alpha\mu}_k\;, \label{e1}\\ \theta_{k 1}^{\alpha\mu}&:=& \sqrt{\frac{\hbar}{2}} (\zeta_k^{\alpha\mu}+ \zeta_k^{\alpha\mu\dagger})\;, \label{e2} \\ \theta_{k 0}^{\alpha\mu}&:=& -i\sqrt{\frac{\hbar}{2}} (\zeta_k^{\alpha\mu}-\zeta_k^{\alpha\mu\dagger})\;, \label{e3}\\ [\hspace{-.3mm}[ \theta_{im}^{\alpha\mu},\theta_{jn}^{\beta\nu}]\hspace{-.3mm}]&:=& \hbar\delta_{ij}\delta^{\alpha\beta} [i(1-\mu)(1-\nu)\epsilon_{mn}+\mu\nu\delta_{mn}]\;, \label{e4} \end{eqnarray} where $\zeta$'a are the Green components of $a$'s \cite{green}, $\alpha, \beta,\mu,\nu=0,1$, $m,n=1,2$, and $[\hspace{-.3mm}[~,~]\hspace{-.3mm}]$ is the parabraket: \begin{equation} [\hspace{-.3mm}[ \theta_{im}^{\alpha\mu},\theta_{jn}^{\beta\nu}]\hspace{-.3mm}] := \theta_{im}^{\alpha\mu}\theta_{jn}^{\beta\nu} -(-1)^{\mu\nu+\alpha+\beta} \theta_{jn}^{\beta\nu}\theta_{im}^{\alpha\mu}\;, \label{pbraket} \end{equation} introduced in \cite{p10}. Note that Eq.~(\ref{e4}) is the statement of the canonical quantization rule for the Green components $\theta$ on the one hand, and the expression of the normal relative statistics \cite{g-m,p10} on the other. The classical analogs of the self-adjoint operators $\theta$ are obtained by setting $\hbar=0$ in Eq.~(\ref{e4}). One also generalizes the definition of the parabracket to arbitrary polynomials in $\theta$'s, according to: \begin{equation} [\hspace{-.3mm}[ M,N ]\hspace{-.3mm}]=MN-(-)^{\eta(M,N)}MN\;, \label{e5} \end{equation} where $M$ and $N$ are monomials: \begin{eqnarray} M&:=&\theta_{i_1 m_1}^{\alpha_1\mu_1}\cdots\theta_{i_r m_r}^{ \alpha_r \mu_r}\;,\nonumber\\ N&:=&\theta_{j_1 n_1}^{\beta_1\nu_1}\cdots\theta_{j_s n_s}^{ \beta_s \nu_s}\;,\nonumber\\ \eta(M,N)&:=&(\sum_{k=1}^r\mu_k)(\sum_{l=1}^s\nu_l)+ r\sum_{l=1}^s\beta_l+s\sum_{k=1}^r\alpha_k\;, \label{e6} \end{eqnarray} and bilinearity of the parabracket. In the classical limit the parabracket of any two polynomials vanishes identically. In Ref.~\cite{p10}, it is also argued that in the Lagrangian formulation of the para-classical mechanics, the Green components of the $\pi b$ coordinate variables are $\theta^{\alpha \mu=0}_{i m=1}$. Thus, one introduces a collective index $I =(i,m)$ which may take $(i=1,\cdots, n_{\pi b};m=1)$ for $\mu=0$ and $(i=1,\cdots,n_{\pi f};m=1,2)$ for $\mu=1$, and denote the Green components of the coordinate variables by $\theta_I^{\alpha\mu}$. The physical quantities, such as a Lagrangian, is chosen from the algebra of polynomials in the coordinates \begin{equation} \psi^\mu_I:=\sum_{\alpha=0}^1\theta^{\alpha\mu}_I \label{e7} \end{equation} and the velocities $\dot{\psi}^\alpha_I$. For computational convenience, they are then expressed in terms of the Green components $\theta^{\alpha\mu }_I$ and $\dot{\theta}^{\alpha\mu}_I$. As a polynomial in (the classical) $\theta$'s and $\dot{\theta}$'s, a Lagrangian must satisfy (up to total time derivatives) the following conditions \cite{p10}: \begin{itemize} \item[1)] It must be real. \item[2)] It must be an even polynomial in both $\pi b$ ($\mu=0$) and $\pi f$ ($\mu=1)$ variables. \end{itemize} To define the notion of reality in the algebra of polynomials in $\theta$'s and $\dot\theta$'s (alternatively in $\psi$'s and $\dot\psi$'s), one first introduces a $*$--operation satisfying: $$(\xi_{x_1}\cdots\xi_{x_n})^*=\xi_{x_n}\cdots\xi_{x_1}\;,$$ $$(\lambda_1P_1+\lambda_2P_2)^*=\lambda_1^*P_1^*+\lambda_2^*P_2^*\;,$$ where $\xi_{x_i}$ are any of the generators: $\theta$'s and $\dot\theta$'s (resp.\ $\psi$'s and $\dot\psi$'s), $\lambda_a\in\relax\hbox{\kern.25em$\inbar\kern-.3em{\rm C}$}$ with $a=1,2$, $\lambda_a^*$ are their complex conjugates, and $P_a$ are polynomials in $\xi_{x_i}$. Then a polynomial $P$ is defined to be real if $P^*=P$. The classical dynamics of the system is given by the least action principle, where the action functional has the form: $S=\int L dt$. This leads to the analogs of the Euler-Lagrange equations: \begin{equation} \frac{d}{dt}(L\frac{\stackrel{\leftarrow}{\partial}}{\partial \dot{\theta}})- L\frac{\stackrel{\leftarrow}{\partial}}{\partial\theta}=0\;. \label{e8} \end{equation} Here the indices on $\theta$'s are suppressed for simplicity and the left partial derivatives with respet to $\theta$'s and $\dot\theta$'s are defined in Refs.~\cite{o-k,p10}. Having reviewed the basic elements of the Lagrangian formulation of para-classical systems, we would like to address the question: \begin{itemize} \item[] {\em Does the algebraic structure of ($p=2$) parastatistical dynamical variables allow for a transformation of $\pi b$ variables into $\pi f$ variables and vice versa?} \end{itemize} Unlike, the case of ordinary ($p=1$) fermi -- bose systems, where the product of two fermi variables is a commutative algebraic object and thus behaves as a bose variable, the algebraic structure of the ($p=2$) variables is too complicated to have such a simple grading. Nevertheless, in view of the formalism developed in \cite{p10}, one can easily respond to the above mentioned question in the positive. To see this, consider the algebra $B$ of the real Green components generated by $\xi_k^{\alpha\mu}$, and the algebra $A$ generated by: $$\gamma_k^\mu=\sum_{\alpha=0}^1 \xi_k^{\alpha\mu}\;.$$ The elements of $A$ (resp.\ $B$) will be used as non-dynamical parameters added to the algebra of polynomials in dynamical variables $\psi$'s and $\dot\psi$'s (resp.\ $\theta$'s and $\dot\theta$'s). Then in the enlarged algebra, it is not difficult to check that the multiplication of dynamical variables $\psi^\mu_I$ and $\dot\psi^\mu_I$ by the real parameters: \begin{equation} \gamma_k=\{\gamma^0_k,\gamma^1_k\}:= \sum_{\alpha,\beta=0}^1\{\xi^{\alpha \mu=0}_k, \xi^{\beta \mu=1}_k\}\;, \label{e9} \end{equation} changes their parity. Here there is no summation over the index $k$. This can be easily verified by defining $\psi^{\alpha\mu'}_{I}:=\psi^{\alpha\mu}_I \gamma_k$ and examining their commutation properties by first decomposing them into their Green components. One can further show that $\gamma_k$ commute with all the parabose variables and anticommute with all the parafermi variables. Presence of $\gamma_k$ allows for the existence of the $\pi SUSY$ transformations. We shall examine examples of such symmetry transformations in the next section. We shall also introduce $\delta\gamma_k$ which are analogs of the (fermionic) parameters of the infinitesimal supersymmetry transformations. \section{SUSY and $\pi$SUSY Oscillators} A thorough discussion of the supersymmetric (SUSY) oscillator is offered in Ref.~\cite{bd}. The Hamiltonian operator of one-dimensional SUSY oscillator is the sum of the Hamiltonians of a fermi and a bose oscillators with identical frequencies, i.e., \begin{eqnarray} \hat{H}&=&\hat{H}^0+\hat{H}^1\;, \label{e10}\\ \hat{H}^0&:=&\frac{\omega}{2}\{\hat a^\dagger,\hat a\}\;, \label{e10.1}\\ \hat H^1&:=&\frac{\omega}{2}[\hat \alpha^\dagger,\hat \alpha]\;. \label{e10.2} \end{eqnarray} Here, $\hat a$ and $\hat a^\dagger$ (resp.\ $\hat\alpha$ and $\hat\alpha^\dagger$) stand for the bosonic (resp.\ fermionic) annihilation and creation operators and the hats are placed to distinguish the quantum mechanical operators and the classical dynamical variables. The combined system of two oscillators (\ref{e10}) serves as a simple example of a supersymmetric system. To reveal the supersymmetry of this system, we shall first switch to the self-adjoint operators: \begin{eqnarray} \hat x&:=&\sqrt{\frac{\hbar}{2\omega}}\,(\hat a+\hat a^\dagger)\;, \nonumber\\ \hat p&:=&-i\sqrt{\frac{\omega\hbar}{2}}\,(\hat a -\hat a^\dagger)\;, \label{e11}\\ \hat \psi_1&=&\sqrt{\frac{\hbar}{2}}\,(\hat\alpha+\hat\alpha^\dagger)\;, \nonumber\\ \hat \psi_2&=&-i\sqrt{\frac{\hbar}{2}}\,(\hat\alpha-\hat\alpha^\dagger)\;.\nonumber \end{eqnarray} Then the Hamiltonian (\ref{e10}) takes the form: \begin{equation} \hat H=\frac{1}{2}(\hat p^2+\omega^2 \hat x^2)+\frac{i\omega}{2} \epsilon_{mn}\hat \psi_m\hat \psi_n\;, \label{e12} \end{equation} where $\epsilon_{mn}$ are the components of the Levi Civita symbol. The classical counterpart of the SUSY oscillator is obtained by dropping the hats in the above relations and treating $x$ and $p$ as bosonic (commuting or even) and $\psi_m$ as fermionic (anticommuting or odd) supernumbers \cite{bd}, respectively. The classical SUSY oscillator may also be described using the Lagrangian: \begin{equation} L=\frac{1}{2}(\dot{x}^2-\omega^2x^2)+\frac{i}{2}\delta_{mn} (\psi_m\dot \psi_n-\dot \psi_m\psi_n)-\frac{i\omega}{2}\epsilon_{mn}\psi_m \psi_n\;, \label{e13} \end{equation} where $m,n=1,2$. Then it is an easy exercise to check that this Lagrangian is invariant (up to total time derivatives) under the transformation: \begin{eqnarray} \delta x&=&i\psi_m\,\delta\zeta_m \;, \label{e14}\\ \delta \psi_m&=&(\delta_{mn}\dot x+\omega\epsilon_{mn} x)\,\delta\zeta_n\;, \label{e15} \end{eqnarray} where $\delta\zeta_n$ are ``infinitesimal'' fermionic supernumber parameters \cite{bd}. The corresponding N\"other charges of this symmetry -- the supercharges -- are given by: \begin{equation} Q_m=\lambda (\delta_{mn}\dot x-\omega\epsilon_{mn}x)\psi_n\;, \label{charge} \end{equation} where $\lambda\in\relax\hbox{\kern.25em$\inbar\kern-.3em{\rm C}$}$ is an arbitrary non-zero coefficient. Upon quantization of this system one can easily show that the supercharges, that generate the supersymmetry transformations, and the Hamiltonian satisfy the defining algebra of SQM. In particular, taking $\lambda=1/\sqrt{\hbar}$, one has: \begin{equation} \{ \hat Q_m,\hat Q_n\}=2\delta_{mn}\hat H\;. \label{sqm} \end{equation} Next, let us introduce the para-generalization of the SUSY oscillator. We shall denote this by $\pi$SUSY oscillator for simplicity. In general, the Hamiltonian for the parabose and parafermi oscillators is given by Eqs.~(\ref{e10.1}) and (\ref{e10.2}), with $\hat a$ and $\hat\alpha$, now, denoting the parabose and parafermi annihilation operators, respectively, \cite{o-k}. Returning to our notation of Sec.~2, we set $\hat a:=\hat a^{\mu=0}$ and $\hat\alpha:=\hat a^{\mu=1}$. In terms of the self-adjoint operators: $\hat\psi^\mu:=\hat\psi^\mu_{i=1}$ of Eq.~(\ref{e7}) and their Green components $\theta_m^{\alpha\mu}:=\theta_{i=1,m}^{\alpha\mu}$, we have: \begin{eqnarray} \hat H^\mu&=&\frac{\omega}{2}[ (1-\mu)\delta_{mn}\hat\psi^\mu_m \hat\psi^\mu_n+i\mu\epsilon_{mn}\hat\psi_m^\mu\hat\psi_n^\mu ]\;, \label{e16}\\ &=&\frac{\omega}{2}[ (1-\mu)\delta_{mn} \hat\theta_m^{\alpha\mu}\hat\theta_n^{\alpha\mu}+ i\mu\epsilon_{mn}\hat\theta_m^{\alpha\mu}\hat\theta_n^{\alpha\mu}]\;, \label{e17} \end{eqnarray} where $\mu=0,1$ correspond to $\pi b$ and $\pi f$ oscillators, respectively. The ($p=2$) -- $\pi$SUSY oscillator is then defined by Eq.~(\ref{e10}): \begin{equation} \hat H=\hat H^0+\hat H^1=\sum_{\alpha=0}^1\left\{ [\frac{1}{2}(\hat\pi^\alpha)^2+\frac{\omega^2}{2}(\hat\chi^\alpha)^2] +\frac{i\omega}{2}\epsilon_{IJ}\,\hat\tau^\alpha_I\hat\tau^\alpha_J \right\}\;, \label{e18} \end{equation} where $\hat\chi^\alpha:=\hat\theta_1^{\alpha\mu=0}/\sqrt{\omega},~ \hat\pi^\alpha:=\sqrt{\omega}\hat\theta_2^{\alpha\mu=0}$ are the Green components of the $\pi b$ coordinate and momentum operators, and $\hat\tau^\alpha_I:=\hat\theta^{\alpha\mu=1}_{m=I}$ are those of the $\pi f$ coordinate operators. The Lagrangian associated with the $\pi$SUSY oscillator is given by: \begin{eqnarray} L&=&\frac{1}{2}(\dot x^2-\omega^2x^2)+\frac{i}{4}\delta_{IJ} (\psi_I\dot\psi_J-\dot\psi_I\psi_J)-\frac{i\omega}{2} \epsilon_{IJ}\,\psi_I\psi_J\;, \label{e20}\\ &=&\sum_{\alpha=0}^1\left\{ \frac{1}{2}[(\dot \chi^\alpha)^2- \omega^2(\chi^\alpha)^2]+\frac{i}{2}\delta_{IJ}\tau^\alpha_I \dot\tau_J^\alpha-\frac{i\omega}{2}\epsilon_{IJ}\,\tau^\alpha_I \tau_J^\alpha\right\}\;, \label{e21} \end{eqnarray} where $x=\sum_{\alpha=0}^1\chi^\alpha,~\psi_I=\sum_{\alpha=0}^1\tau_I^\alpha$ are the ($p=2$) $\pi b$ and $\pi f$ dynamical variables, respectively. The form of the $\pi b$ and $\pi f$ kinetic terms in (\ref{e20}) is obtained in Ref.~\cite{p10} in an attempt to consistently generalize the Peierls bracket quantization scheme to the paraclassical systems. The Peierls bracket quantization of this system leads to the following paracommutation relations: \begin{eqnarray} [\hspace{-.3mm}[ \hat\chi^\alpha,\hat\chi^\beta]\hspace{-.3mm}]&=&0\;,\nonumber\\ [\hspace{-.3mm}[\hat\chi^\alpha,\hat{\dot{\chi}}^\beta]\hspace{-.3mm}]&=&i\hbar \delta^{\alpha\beta}\;,\nonumber\\ [\hspace{-.3mm}[\hat{\dot\chi}^\alpha,\hat{\dot\chi}^\beta]\hspace{-.3mm}]&=&0\;, \label{e30}\\ [\hspace{-.3mm}[\hat\tau^\alpha_I,\hat\tau^\beta_J]\hspace{-.3mm}]&=&\hbar \delta^{\alpha\beta}\delta_{IJ}\;,\nonumber \\ [\hspace{-.3mm}[\hat\chi^\alpha,\hat\tau_I^\beta]\hspace{-.3mm}]&=&0\;,\nonumber \\ [\hspace{-.3mm}[ \hat{\dot{\chi}}^\alpha,\hat\tau_I^\beta]\hspace{-.3mm}]&=&0\;,\nonumber \end{eqnarray} which become identical with the canonical quantization relations (\ref{e4}) if one only considers the momenta $\pi^\alpha$ conjugate to $\chi^\alpha$ and identifies them with $\dot{\chi}^\alpha$. \section{Symmetries of the $\pi$SUSY Oscillator} Setting $\tau^0_I=\tau^1_I$ and $\chi^0=\chi^1$ in (\ref{e21}), one recovers the Lagrangian for the SUSY oscillator (\ref{e13}). This may be used as a hint to seek similar symmetries for the $\pi$SUSY oscillator. Following this hint, consider the $\pi$-SUSY transformation: \begin{eqnarray} \delta\chi^\alpha&=&i\tau^\alpha_J\,\delta\gamma_J\;, \label{e31}\\ \delta\tau_I^\alpha&=& (\delta_{IJ}\,\dot\chi^\alpha+\omega \epsilon_{IJ}\,\chi^\alpha)\,\delta\gamma_J\;, \label{e32} \end{eqnarray} where $\delta\gamma_J$ are the ``infinitesimal'' analogs of $\gamma_J$ of Eq.~(\ref{e9}). It is not difficult to check that the action functional and therefore the dynamical equations remain invariant under this transformation. Indeed, one finds: \begin{equation} \delta L \propto \frac{d}{dt}( \dot\chi^\alpha\tau^\alpha_J - \omega\epsilon_{JI}\tau_I^\alpha\chi^\alpha)\,\delta\gamma_J\;. \label{e33} \end{equation} Thus the corresponding conserved charges have the form: \begin{equation} Q^1_J=\lambda( \dot\chi^\alpha\tau^\alpha_J - \omega\epsilon_{JI}\tau_I^\alpha \chi^\alpha)\;. \label{e34} \end{equation} Here the superscript ``1'' is placed for later use and $\lambda$ is a non-zero numerical coefficient. In the remainder of this paper, we shall set $\hbar=1$ for simplicity. The quantum analog of $Q^1_J$ with an appropriate normalization is given by: \begin{equation} \hat Q^1_J:=\hat{\dot\chi}^\alpha \hat\tau^\alpha-\omega\epsilon_{JI}\hat\chi^\alpha \hat\tau_I^\alpha\;. \label{e35} \end{equation} In view of the paracommutation relations (\ref{e30}), it is not difficult to check that $Q_J^1$ generate the transformations (\ref{e31}) and (\ref{e32}), i.e., \begin{eqnarray} [\hspace{-.3mm}[ \hat\chi^\alpha,\hat Q_J^1\delta\gamma_J]\hspace{-.3mm}]&=& i\hat\tau^\alpha_J\delta\gamma_J\:=\: \delta\hat\chi^\alpha\;, \label{e36}\\ [\hspace{-.3mm}[ \hat\tau^\alpha_I,\hat Q_J^1\delta\gamma_J]\hspace{-.3mm}]&=& (\delta_{IJ}\hat{\dot{\chi}}^\alpha+\omega \epsilon_{IJ}\hat\chi^\alpha)\,\delta\gamma_J\:=\: \delta\hat\tau^\alpha_I\;, \label{e37} \end{eqnarray} and that they satisfy the defining algebra of SQM, namely: \begin{equation} \{ \hat Q_I^1,\hat Q_J^1\}=2\delta_{IJ}\hat H\;. \label{e38} \end{equation} Note also that $\hat Q^1_I$ are self-adjoint operators by construction (\ref{e35}). Another important point in handling ($p=2$) para-dynamical systems is that the Green components are not the physical dynamical variables. In other words, one must be able to express all physical quantities in terms of the variables, $x,~\dot x,~\psi_I,$ and $\dot\psi_I$. This also applies to the $Q_I^1$. In fact, one can show that: \begin{equation} \hat Q_J^1=\frac{1}{2}\{ \delta_{JK}\hat{\dot x}-\omega\epsilon_{JK}\hat x\:,\:\hat\psi_K\}\;. \label{e39} \end{equation} Here use is made of the identities: \begin{equation} \hat{\dot\chi}^\alpha \hat\tau^\alpha_I=\frac{1}{2}\{\hat{\dot x}, \hat\psi\}\;,~~~~~~ \hat\chi^\alpha \hat\tau^\alpha_I=\frac{1}{2} \{\hat x,\hat\psi_I\}\;. \label{e40} \end{equation} The $\pi$SUSY transformations (\ref{e31}) and (\ref{e32}) mix the Green components $\chi^\alpha$ and $\tau_I^\alpha$ with the same Green index $\alpha$. Since the Green components are not physical quantities, there must be no difference between say $\tau^0_I$ and $\tau^1_I$. This suggests the possibility of symmetry transformations which mix $\chi^\alpha$ with $\tau^{\alpha+1}$. Here the values of the Green indices is taken in $\ Z \hspace{-.08in}Z_2$, i.e., they are calculated modulo $2$. The following is such a symmetry transformation: \begin{eqnarray} \delta\chi^\alpha&=&-i\tau^{\alpha+1}_J\,\delta\gamma_J\;, \label{e41}\\ \delta\tau_I^\alpha&=& (\delta_{IJ}\dot\chi^{\alpha+1}+\omega \epsilon_{IJ}\chi^{\alpha+1})\,\delta\gamma_J\;. \label{e42} \end{eqnarray} The associated conserved charges to this symmetry are given by \begin{equation} Q^2_J=\lambda'( \tau^\alpha_J\dot\chi^{\alpha+1}- \omega\epsilon_{JK}\tau_K^\alpha\chi^{\alpha+1})\;, \label{e43} \end{equation} where the summation over $\alpha$ is understood. Quantizing the system and taking: \begin{equation} \hat Q^2_J:=i(\hat\tau^\alpha_J\hat{\dot{\chi}}^{\alpha+1}- \omega\epsilon_{JK}\hat\tau_K^\alpha\hat\chi^{\alpha+1})\;, \label{e44} \end{equation} one obtains another set of self-adjoint $\pi$SUSY charges. They generate the transformations (\ref{e41}) and (\ref{e42}) and are expressed in terms of the physical variables $x$ and $\psi$ according to: \begin{equation} \hat Q^2_J= \frac{-i}{2}[\delta_{JK}\hat{\dot x}-\omega\epsilon_{JK}\hat x\:,\: \hat\psi_K]\;. \label{e45} \end{equation} Here use is made of the identities: \begin{equation} \hat\tau_I^\alpha\hat{\dot\chi}^{\alpha+1}=\frac{1}{2} [\hat\psi_I,\hat{\dot x}]\;,~~~~~~ \hat\tau_I^\alpha\hat\chi^{\alpha+1}=\frac{1}{2}[\hat\psi_I,\hat x]\;. \label{e46} \end{equation} Furthermore, the superalgebra relation: \begin{equation} \{ \hat Q^2_I,\hat Q^2_J\}=2\delta_{IJ}\hat H\;, \label{e47} \end{equation} also holds. The next natural step in the study of the symmetries of the $\pi$SUSY oscillator is to investigate the algebraic properties of both types of $\pi$SUSY's. Proceeding in this direction, one finds: \begin{equation} [\hspace{-.3mm}[ \hat Q_J^a,\hat Q_K^b]\hspace{-.3mm}]= \{ \hat Q_J^a,\hat Q_K^b\}=2\delta_{JK}\delta^{ab}\hat H -2\epsilon^{ab}\epsilon_{JK}\hat {\cal Q}\;, \label{e48} \end{equation} where \begin{equation} \hat{\cal Q}:=i\omega\hat\chi^\alpha\hat{\dot\chi}^{\alpha+1}+ \frac{\omega}{2}\hat\tau_I^\alpha\hat\tau_I^{\alpha+1} \label{e49} \end{equation} is another (self-adjoint) conserved charge. Repeating this procedure, i.e., including ${\cal Q}$ in the set of the generators of symmetries and investigating the parabracket of ${\cal Q}$ and other generators, one obtaines \begin{equation} [\hspace{-.3mm}[ \hat Q^a_J,\hat{\cal Q}]\hspace{-.3mm}]=[ \hat Q^a_J,\hat{\cal Q}]=0\;. \label{e50} \end{equation} Thus the {\em superalgebra} consisting of the generators of $\pi$SUSY of the $\pi$SUSY oscillator closes. Summarizing the superalgebra relations, one has: \begin{eqnarray} [\hspace{-.3mm}[ \hat Q_J^a,\hat H]\hspace{-.3mm}]&=&[\hat Q_J^a,\hat H]\:=\:0\nonumber\\ [\hspace{-.3mm}[ \hat Q_J^a,\hat Q_K^b]\hspace{-.3mm}]&=& \{ \hat Q_J^a,\hat Q_K^b\}\: =\: 2\delta_{JK}\delta^{ab}\hat H -2\epsilon^{ab}\epsilon_{JK}\hat {\cal Q}\;,\nonumber\\ [\hspace{-.3mm}[ \hat{\cal Q},\hat H]\hspace{-.3mm}]&=&[ \hat{\cal Q},\hat H]\:=\: 0 \label{e60}\\ [\hspace{-.3mm}[ \hat Q^a_J,\hat{\cal Q}]\hspace{-.3mm}]&=&[ \hat Q^a_J,\hat{\cal Q}] \:=\: 0\;.\nonumber \end{eqnarray} The generators $Q_J^a$ behave as the ``odd'' elements of the super Lie algebra and $H$ and ${\cal Q}$ as the ``even'' (central) elements. The (central) charge ${\cal Q}$ is also expressed in terms of the physical variables. One has: $$ {\cal Q}=\frac{i\omega}{2}(\hat x\hat{\dot x}-\hat{\dot x}\hat x) +\frac{\omega}{2}\delta_{IJ}\hat\psi_I\hat\psi_J\;. $$ Here, one uses the following identities: $$\delta_{IJ}\hat\psi_I\hat\psi_J= \hat\tau^\alpha_I\hat\tau^{\alpha+1}_I+2\;,~~~~~~ \hat x\hat{\dot x}-\hat{\dot x}\hat x= 2\hat\chi^\alpha\hat{\dot\chi}^{\alpha+1}+2i\;.$$ One can also examine the symmetry transformations generated by ${\cal Q}$. These are obtained by computing: \begin{eqnarray} [\hspace{-.3mm}[ \hat\chi^\alpha,{\cal Q}\delta\epsilon]\hspace{-.3mm}]&=& \{\hat\chi^\alpha,{\cal Q}\} \delta\epsilon\:=\:\omega \hat\chi^{\alpha+1}\delta\epsilon\;,\nonumber\\ [\hspace{-.3mm}[ \hat\tau^\alpha_I,{\cal Q}\delta\epsilon]\hspace{-.3mm}]&=& \{\hat\tau^\alpha_I,{\cal Q}\}\delta\epsilon\:=\:\omega \hat\tau^{\alpha+1}_I\delta\epsilon\;.\nonumber \end{eqnarray} Thus: $$\delta_{\cal Q}\chi^\alpha=\omega\, \chi^{\alpha+1}\delta\epsilon\;,~~~~~~ \delta_{\cal Q}\tau^\alpha_I=\omega\, \tau^{\alpha+1}_I\delta\epsilon\;.$$ Here $\delta\epsilon$ is an infinitesimal commuting parameter. In terms of the physical dynamical variables, one has: $$\delta_{\cal Q}x=\omega\,x\,\delta\epsilon\,\;,~~~~~~ \delta_{\cal Q}\psi_I= \omega\,\psi_I\,\delta\epsilon\,\;.$$ We would like to conclude this section by emphasizing the enormous advantage of using parabracket (\ref{pbraket}) in performing the tedius computations necessary for establishing the superalgebra relations Eqs.~(\ref{e60}). The details of these computations have been omitted due to the space limitations. \section{Degeneracy Structure of General SPQM} Let us first define SPQM: \begin{itemize} \item[] {\bf Definition:} {\em Let ${\cal H}$ be a $\ Z \hspace{-.08in}Z_2$-graded Hilbert space with grading involution {\large $\hat\tau$}. Then a quantum mechanical system with ${\cal H}$ as the Hilbert space and self-adjoint symmetry generators $\hat Q_{I_n}^{a_n}$, $\hat{\cal Q}_n$, $n=1,\cdots N$, ~ $I_n,a_n=1,2$, and the Hamiltonian operator $\hat H$ satisfying the super Lie algebra relations: \begin{eqnarray} [\hat Q_{J_n}^{a_n},\hat H]&=&[\hat{\cal Q}_n,H]\:=\: [ \hat Q^{a_n}_{J_n},\hat{\cal Q}]\:=\:0\label{e70}\\ \{ \hat Q_{J_n}^{a_n},\hat Q_{K_m}^{b_m}\}&=&\delta_{nm}( 2\delta_{J_nK_n}\delta^{a_nb_n}\hat H -2\epsilon^{a_nb_n}\epsilon_{J_nK_n}\hat{\cal Q}_n)\;, \label{e71} \end{eqnarray} and parity properties: \begin{equation} \{\mbox{\large$\hat\tau$},\hat Q_{I_n}^{a_n}\}=0\;,~~~~~ [\mbox{\large$\hat\tau$},\hat{\cal Q}_n]= [\mbox{\large$\hat\tau$},\hat H]=0\;, \label{e72} \end{equation} for all $I_n,a_n$ and $n=1,\cdots,N$, is called a ($p=2$) -- supersymmetric paraquantum mechanical (SPQM) system of type $N$.} \end{itemize} In this section, we shall present a detailed analysis of the spectrum degeneracy structure of general ($p=1$)-SPQM systems of type $N=1$. For $N=1$ we suppress the index $n=1$ and recover the super Lie algebra of the $\pi$SUSY oscillator, i.e., Eqs.\ (\ref{e60}). For simlicity we shall drop the hats and introduce the notation: $$Q_1\equiv Q_1^1\;,~~~Q_2\equiv Q_2^1\;,~~~Q_3\equiv Q_1^2\;, ~~~Q_4\equiv Q_2^2\;.$$ Then Eqs.~(\ref{e60}) are written as: \begin{eqnarray} Q_i^2&=&H\;,\label{e81}\\ \{ Q_1,Q_2\}&=&0\;,\label{e82}\\ \{ Q_1,Q_3\}&=&0\;,\label{e83}\\ \{ Q_1,Q_4\}&=&-2 {\cal Q}\;,\label{e85}\\ \{ Q_2,Q_3\}&=&2 {\cal Q} \;,\label{e84}\\ \{ Q_2,Q_4\}&=&0\;,\label{e86}\\ \{ Q_3,Q_4\}&=&0\;,\label{e87}\\ {[} Q_i,{\cal Q}{]}&=&0\;,\label{e88} \end{eqnarray} where $i=1,2,3,4$. Next, we use the simultaneous eigenstate vectors $|E,q_1,q\rangle$, with $E,q_1,q\in\relax{\rm I\kern-.18em R}$, of $H,~ Q_1$ and ${\cal Q}$ to span the Hilbert space. We shall assume that these state vectors form an orthonormal basis and attempt to represent all the relevant operators in this basis. These properties are summarized by the following set of relations: \begin{eqnarray} H|E,q_1,q\rangle&=&E|E,q_1,q\rangle\;,~~~~~~ Q_1|E,q_1,q\rangle\:=\: q_1|E,q_1,q\rangle\;, \label{e91}\\ {\cal Q}|E,q_1,q\rangle&=&q|E,q_1,q\rangle\;,~~~~~~ \langle E',q'_1,q'|E,q_1,q\rangle\:=\: \delta_{E'E}\delta_{q'_1,q_1}\delta_{q'q}\;. \label{e92} \end{eqnarray} A simple consequence of Eq.~(\ref{e81}) with $i=1$, is that the energy spectrum is non-negative. Furthermore, for any energy level $E$, one has: \begin{eqnarray} q_1&=&\pm\sqrt{E}\;, \label{e101}\\ |q_1,q\rangle=0&\Leftrightarrow&|-q_1,q\rangle=0\;, \label{e102}\\ Q_2|q_1,q\rangle&=&C_2(q_1,q)|-q_1,q\rangle\;,~~~~~~C_2(q_1,q)\in\relax\hbox{\kern.25em$\inbar\kern-.3em{\rm C}$}-\{0\}\;, \label{e109}\\ Q_3|q_1,q\rangle&=&C_3(q_1,q)|-q_1,q\rangle\;,~~~~~~C_3(q_1,q)\in\relax\hbox{\kern.25em$\inbar\kern-.3em{\rm C}$}-\{0\}\;, \label{e110} \end{eqnarray} where use is made of Eqs.~(\ref{e81}) -- (\ref{e83}) and abbreviation $|q_1,q\rangle$ is used for $|E,q_1,q\rangle$. Enforcing Eq.~(\ref{e84}), one finds: \begin{equation} C_2(q_1,q)C_3(-q_1,q)+C_3(q_1,q)C_2(-q_1,q)=2 q\;. \label{e111} \end{equation} Then by acting both sides of Eqs.~(\ref{e81}), with $i=2,3$, on $|q_1,q\rangle$, one has: \begin{equation} C_2(q_1,q)C_2(-q_1,q)=E\;,~~~~~~C_3(q_1,q)C_3(-q_1,q)=E\;. \label{e115} \end{equation} Next, we calculate: $$E=(\langle q_1,q|Q_2)(Q_2|q_1,q\rangle)=C_2(q_1,q)^*C_2(q_1,q)\;.$$ A similar relation holds for $C_3$. These relations together with Eqs.~(\ref{e115}) imply: \begin{equation} C_2(\pm q_1,q)=\sqrt{E}\,e^{\pm i\alpha_2(q)}\;,~~~~~~ C_3(\pm q_1,q)\:=\:\sqrt{E}\,e^{\pm i\alpha_3(q)}\;. \label{e116} \end{equation} Combining the latter equations with Eq.~(\ref{e111}), one is led to: \begin{equation} \frac{C_2(q_1,q)}{C_3(q_1,q)}+\frac{C_3(q_1,q)}{C_2(q_1,q)} =\frac{2q}{E}\;. \label{e117} \end{equation} Eqs.~(\ref{e116}) and (\ref{e117}), in turn, yield: \begin{equation} \cos[\alpha_2(q)-\alpha_3(q)]=\frac{q}{E}\;. \label{e121} \end{equation} Next, we act both sides of Eqs.~(\ref{e85}), (\ref{e86}), and (\ref{e87}) on $|q_1,q\rangle$ on the left. This gives rise to: \begin{eqnarray} Q_1Q_4|q_1,q\rangle&=&-q_1Q_4|q_1,q\rangle-2 q|q_1,q\rangle\;, \label{e112}\\ Q_2Q_4|q_1,q\rangle&=&-\sqrt{E}\,e^{i\alpha_2}Q_4|-q_1,q\rangle\;, \label{e119}\\ Q_3Q_4|q_1,q\rangle&=&-\sqrt{E}\,e^{i\alpha_3}Q_4|-q_1,q\rangle\;. \label{e120} \end{eqnarray} To pursue our analysis further, we express the action of $Q_4$ on the basic kets $|q_1,q\rangle$ as the following linear combination: \begin{equation} Q_4|q_1,q\rangle=: a(q_1,q)|\sqrt{E},q\rangle+b(q_1,q)|-\sqrt{E},q\rangle \;. \label{e130} \end{equation} where $a$ and $b$ are complex numbers {\em a priori} depending on $q_1$, $q$ and of course $E$. Substituting this expression in Eq.~(\ref{e112}), one finds: \begin{equation} a(q_1=\sqrt{E},q)=-\frac{q}{\sqrt{E}}\,~~~~~ a(q_1=-\sqrt{E},q)=0\;. \label{e131} \end{equation} Repeating the same procedure for Eqs.~(\ref{e119}) and (\ref{e120}), and performing the simple algebra, one finally obtains: \begin{eqnarray} b(q_1=\sqrt{E},q)&=&0\;,~~~~~b(q_1=-\sqrt{E},q)\:=\:-a(q_1= \sqrt{E},q)=\frac{q}{\sqrt{E}}\;, \label{e6.1}\\ e^{i\alpha_2(q)}\:=\:\pm 1 &=&e^{i\alpha_3(q)}\;. \label{e6.2} \end{eqnarray} The last pair of equations together with Eq.~(\ref{e121}) imply: \begin{equation} q=E\eta\;,~~~~~~~~\eta=\pm 1\;. \label{eq} \end{equation} Having obtained all the unknowns of our construction and appealing to the gauge freedom of the phases of the initial basic eigenstate vectors -- which allows us to set, say, $\alpha_2=0$ so that $e^{i\alpha_3}=\eta$ -- we are in a position to present matrix reperesentations of all the charges. However, before presenting these representations, we would like to remark that although $|\sqrt{E},q\rangle\neq 0\Leftrightarrow |-\sqrt{E},q\rangle\neq 0$, this relation does not imply that $|\pm\sqrt{E},q\rangle\neq 0$ for some $q$, i.e., in general it may be the case that for some values of $E$ the state vectors corresponding to either $q=+E$ or $q=-E$ vanish. In this case $E$ will be doubly degenerate. Otherwise it will be quadruply degenerate. For the latter case the symmetry generators are represented by: \begin{eqnarray} \left. Q_1\right|_{{\cal H}_E}&=&\sqrt{E}\left(\begin{array}{cccc} 1 & 0 & & \\ 0 & -1 & & \\ & & 1 & 0 \\ & & 0 & -1 \end{array}\right)\nonumber\\ \left. Q_2\right|_{{\cal H}_E}&=& \sqrt{E}\left(\begin{array}{cccc} 0 & 1 & & \\ 1 & 0 & & \\ & & 0 & 1 \\ & & 1 & 0 \end{array}\right)\nonumber\\ \left. Q_3\right|_{{\cal H}_E}&=& \sqrt{E}\left(\begin{array}{cccc} 0 & 1 & & \\ 1 & 0 & & \\ & & 0 & -1 \\ & & -1 & 0 \end{array}\right)\nonumber\\ \left. Q_4 \right|_{{\cal H}_E}&=& \sqrt{E}\left(\begin{array}{cccc} -1 & 0 & & \\ 0 & 1 & & \\ & & 1 & 0 \\ & & 0 & -1 \end{array}\right)\nonumber\\ \left. {\cal Q}\right|_{{\cal H}_E}&=& E\left(\begin{array}{cccc} 1 & 0 & & \\ 0 & 1 & & \\ & & -1 & 0 \\ & & 0 & -1 \end{array}\right)\nonumber\\ \left. H\right|_{{\cal H}_E}&=& E\left(\begin{array}{cccc} 1 & 0 & & \\ 0 & 1 & & \\ & & 1 & 0 \\ & & 0 & 1 \end{array}\right)\;.\nonumber \end{eqnarray} Here we have identified: \begin{eqnarray} |\sqrt{E},E\rangle&=&\left(\begin{array}{c} 1\\0\\0\\0 \end{array}\right)\;,~~~~~~ |-\sqrt{E},E\rangle\:=\:\left(\begin{array}{c} 0\\1\\0\\0 \end{array}\right)\;,\nonumber\\ |\sqrt{E},-E\rangle&=&\left(\begin{array}{c} 0\\0\\1\\0 \end{array}\right)\;,~~~~~~ |-\sqrt{E},-E\rangle\:=\:\left(\begin{array}{c} 0\\0\\0\\1 \end{array}\right)\;,\nonumber \end{eqnarray} the empty blocks consist of vanishing entries, and ${\cal H}_E$ denotes the degeneracy Hilbert space associated with the energy $E>0$. In view of Eqs.~(\ref{e72}), we can also write down the matrix representation of the involution (chirality) operator in this basis. The result is given by $$ \left. \mbox{\large$\tau$}\right|_{{\cal H}_E}= \left(\begin{array}{cccc} 0 & -i\epsilon_1 & & \\ i\epsilon_1 & 0 & & \\ & & 0 &-i\epsilon_2\\ & &i\epsilon_2 & 0 \end{array}\right)\;, $$ where $\epsilon_1,\epsilon_2=\pm 1$. It is an easy exercise to diagonalize the chirality involution(s) and to find out that in the diagonal form it has the form: \begin{equation} \left. \mbox{\large$\tau$}\right|_{{\cal H}_E}= {\rm diag}(1,-1,1,-1)\;. \label{e150} \end{equation} This implies that the quadruply degenerate (positive) energy levels involve two odd (parafermionic) and two even (parabosonic) state vectors. The representations of the symmetry generators and the involution operator for the doubly degenerate energy levels are given by either of the upper-left or lower-right blocks in the above list of matrix representations, according to whether $|\pm\sqrt{E},-E\rangle=0$ or $|\pm\sqrt{E},+E\rangle=0$, respectively. The situation is analogous to the ordinary supersymmetric case, \cite{p8}. The following lemma summarizes our results concerning ($p=2$)-SPQM: \begin{itemize} \item[] {\bf Lemma 1:} {\em The energy spectrum of any ($p=2$) supersymmetric paraquantum system is non-negative. The zero-energy eigenvalue, if exists, is non-degenerate\footnote{This is true provided that other quantum numbers are not present.}. The positive energy levels are either doubly or quadruply degenerate. They consist of pairs of odd and even parity eigenstates.} \end{itemize} Moreover, one can define the Witten index according to $${\rm index_{Witten}}:={\rm trace}(\mbox{\large$\tau$})\;,$$ and use Eq.~(\ref{e150}) to conclude that it counts the difference of the number of even and odd zero-energy states, and that it is a topological invariant. \section{Hilbert Space Structure of the $\pi$SUSY Oscillator} Ref.~\cite{sud} offers an analysis of the energy eigenstates of the one-dimensional parabose oscillator of arbitrary order $p$. In the following, we shall use the results of \cite{sud}, with $p=1$, to construct a complete set of eigenstate vectors for the $\pi$SUSY oscillator. The Hilbert space of the one-dimensional ($p=1$) $\pi$b oscillator is constructed using the following set of orthonormal energy eigenstate vectors: \begin{equation} |n\rangle:=\frac{1}{\sqrt{2^n[\frac{n}{2}]![\frac{n+1}{2}]!}} \: a^{\dagger n}\,|0\rangle\;, \label{e200} \end{equation} where $a^\dagger$ and $|0\rangle$ are the $\pi$b creation operator and the vacuum (ground) state, respectively, and $[k]$ stands for the largest integer smaller than or equal to $k\in\relax{\rm I\kern-.18em R}$. One also has: \begin{eqnarray} H^0|n\rangle&=&(n+1)\omega\;, \label{e201}\\ a|n\rangle&=&\sqrt{2[\frac{n}{2}]+1}\:|n-1\rangle\;, \label{e202}\\ a^\dagger|n\rangle&=&\sqrt{2[\frac{n}{2}]+2}\:|n+1\rangle\;. \label{e203} \end{eqnarray} For the $\pi$SUSY oscillator, one has also the $\pi$f creation and annihilation operators. These have the property that $a^3=0$. So there is an apparent triple grading intrinsic to the ($p=2$) $\pi$f operators. This has been used quite often in the context of parasupersymmetry. In the following we shall demonstrate that this is not the case for the $\pi$SUSY oscillator as one might expect in view of the treatment of Sec.~6. It turns out that the following energy eigenstates form an orthonormal basis for the Hilbert space: \begin{equation} \begin{array}{cc} |n,1\rangle:= \frac{1}{\sqrt{2^n[\frac{n}{2}]![\frac{n+1}{2}]!}}\: a^{\dagger n}|0\rangle\;, ~~~~~~ &(n\geq 0)\\ |n,2\rangle:= \frac{1}{\sqrt{2^n[\frac{n}{2}]![\frac{n-1}{2}]!}} \:\alpha^\dagger a^{\dagger n-1}|0\rangle\;, ~~~~ &(n\geq 1) \\ |n,3\rangle:= \frac{1}{\sqrt{2^n[\frac{n-1}{2}]![\frac{n-2}{2}]!}} \:\alpha^{\dagger 2}a^{\dagger n-2}|0\rangle\;, ~ &(n\geq 2) \\ |n,4\rangle:= \frac{1}{\sqrt{2^n[\frac{n}{2}]![\frac{n-1}{2}]!}}\: a^\dagger\alpha^\dagger a^{\dagger n-2}|0\rangle\; &(n\geq 2) \end{array} \label{e204} \end{equation} Note that the state vector $|n,1\rangle$ is the same as $|n\rangle$ of Eq~(\ref{e200}). To establish the orthonormality of $\{ |n,a\rangle~:~ a=1,2,3,4\}$, one needs to use the following set of paracommutation relations: \begin{eqnarray} \alpha\,\alpha^\dagger\,a^\dagger&=&-a^\dagger \,\alpha^\dagger\, \alpha+2a^\dagger\;, \label{e210}\\ \alpha\,\alpha^{\dagger 2}&=&-\alpha^{\dagger 2}\, \alpha+2\alpha^\dagger\;, \label{q2.4}\\ a\,a^\dagger\,\alpha^\dagger&=&\alpha^\dagger\,a^\dagger\,a+ 2\alpha^\dagger\;, \label{e212}\\ \alpha\, a\, a^\dagger&=&a^\dagger\, a\,\alpha+2\alpha\;, \label{e213}\\ \alpha\,a \, \alpha^\dagger&=&-\alpha^\dagger\, a\,\alpha\;, \label{q23.1} \end{eqnarray} and the identity: \begin{equation} \alpha\, a^{\dagger n}|0\rangle=0\;. \label{e215} \end{equation} Relations (\ref{e210})--(\ref{e215}) are most easily proved in the Green representation. Furthermore, it is not difficult to check that indeed $|n,a\rangle$ are energy eigenvectors, i.e., \begin{equation} H|n,a\rangle=E_n|n,a\rangle\;,~~~~~{\rm with}~~~~~E_n:=n\omega\;. \label{e220} \end{equation} Finally, it is possible to show that $|n,a\rangle$ form a complete set of state vectors. This involves some lengthy algebraic manipulations. The completeness of $\{ |n,a\rangle\}$ results from the following set of relations: \begin{eqnarray} a\,|n,1\rangle&=&\sqrt{2[\frac{n}{2}]+1}\:|n-1,1\rangle\;,\nonumber\\ a^\dagger\,|n,1\rangle&=&\sqrt{2[\frac{n}{2}]+2}\:|n+1,1\rangle\;,\nonumber\\ \alpha\,|n,1\rangle&=&0\nonumber\\ \alpha^\dagger\,|n,1\rangle&=&\sqrt{2}\:|n+1,2\rangle\;,\nonumber\\ a\,|n,2\rangle&=&\sqrt{\frac{[(n-1)/2](2[n/2]-1)}{[n/2]}}\:|n-1,5\rangle\;, \nonumber\\ a^\dagger\,|n,2\rangle&=&\sqrt{2[(n+1)/2]}\: |n+1,5\rangle\;,\nonumber\\ \alpha\,|n,2\rangle&=&\sqrt{2}\,|n-1,1\rangle\;,\nonumber\\ \alpha^\dagger\,|n,2\rangle&=&\sqrt{2}\,|n+1,3\rangle\;,\nonumber\\ a\,|n,3\rangle&=&-\sqrt{2[n/2]}\:|n-1,3\rangle\;,\nonumber\\ a^\dagger\,|n,3\rangle&=&-\sqrt{2[n/2]}\,|n+1,3\rangle\;,\nonumber\\ \alpha\,|n,3\rangle&=&\sqrt{2}\,|n-1,2\rangle\;,\nonumber\\ \alpha^\dagger\,|n,3\rangle&=&0\;,\nonumber\\ a\,|n,4\rangle&=& \sqrt{\frac{[(n-1)/2](2[n/2]-1)+2}{[n/2]}}\:|n-1,2\rangle\;,\nonumber\\ a^\dagger\,|n,4\rangle&=&\sqrt{2[(n+1)/2]}\,|n+1,2\rangle\;,\nonumber\\ \alpha\,|n,4\rangle&=&0\;,\nonumber\\ \alpha^\dagger\,|n,4\rangle&=&0\;,\nonumber \end{eqnarray} where in addition to Eqs.~(\ref{e202})--(\ref{q23.1}), the paracommutation relations: \begin{eqnarray} a\,\alpha^\dagger\,a^\dagger&=&a^\dagger\,\alpha^\dagger\,a\;,\nonumber\\ a\,\alpha^{\dagger 2}&=&-\alpha^{\dagger 2}\,a\;,\nonumber\\ \alpha\, a^\dagger\,\alpha^\dagger&=&-\alpha^\dagger\,a^\dagger\, \alpha\;,\nonumber\\ \alpha^\dagger\, a^\dagger\,\alpha^\dagger&=&0\;,\nonumber \end{eqnarray} are also used. To demonstrate the method of proof of such relations using the Green representation, a proof of the last equation is offered in the following. First note that $$ a=\sum_{\alpha=0}^1\zeta^{\alpha 0}\;,~~~ \alpha=\sum_{\beta=0}^1\zeta^{\beta 1}\;,$$ \begin{equation} [\hspace{-.3mm}[ \zeta^{\alpha \mu},\zeta^{\beta\nu}]\hspace{-.3mm}]=0\;,~~~ [\hspace{-.3mm}[ \zeta^{\alpha \mu},\zeta^{\beta\nu\dagger}]\hspace{-.3mm}]= \delta^{\alpha\beta}\delta^{\mu\nu}\;. \label{e301} \end{equation} Here use is made of Eqs.~(\ref{e1})--(\ref{e4}). Next, one has: \begin{eqnarray} \alpha^\dagger\, a^\dagger\,\alpha^\dagger&=& \sum_{\alpha,\beta,\gamma} \zeta^{\alpha 1\dagger}\zeta^{\beta 0\dagger}\zeta^{\gamma 1 \dagger} \nonumber\\ &=&\sum_{\alpha,\beta,\gamma}(-1)^{1+\beta+\gamma} \zeta^{\beta 0 \dagger}\zeta^{\gamma 1\dagger}\zeta^{\alpha 1 \dagger}\nonumber\\ &=&-\sum_{\alpha,\beta,\gamma} \zeta^{\gamma 1\dagger}\zeta^{\beta 0\dagger}\zeta^{\alpha 1\dagger} \nonumber\\ &=&-\alpha^\dagger\, a^\dagger\,\alpha^\dagger\:=\:0\;, \end{eqnarray} where in the second and third equalities use is made of Eqs.~(\ref{pbraket}) and (\ref{e301}). This concludes our investigation of the energy eigenstates of the $\pi$SUSY oscillator. We summarize the results of this section in the form of the following lemma: \begin{itemize} \item[] {\bf Lemma~2:} {\em The energy spectrum of the $\pi$SUSY oscillator consists of a zero energy non-degenerate ground state (represented by $|n=0,1\rangle$), a doubly degenerate first excited state (level) of energy $E_1=\omega$ (with state vectors $|n=1,1\rangle$ and $|n=1,2\rangle$), and higher excited states of $E_n=n\omega$ ($n\geq 2$) which are quadruply degenerate (with state vectors $|n,a\rangle\;,a=1,2,3,4$.)} \end{itemize} This confirms our general results of Sec.~6. \section{Conclusion} There are dynamical systems involving ($p=2$) parastatistical degrees of freedom and symmetry transformations which mix the parabose and parafermi dynamical variables. The mixing which signifies a parabose -- parafermi supersymmetry is shown to be present because of the non-trivial algebraic properties of such variables. Having established the meaningfullness of the parabose -- parafermi supersymmetry ($\pi$SUSY), one can investigate its relation with the ordinary (bose -- fermi) supersymmetry and the parasupersymmetry. The simple example of an oscillator consisiting of a parabosonic and a parafermionic sector is used to demonstrate the nature of $\pi$SUSY. This oscillator possesses two ordinary supersymmetries. The study of the combined set of generators of these supersymmetries leads to the introduction of a central charge. Thus, it seems that there is no direct relation between parabose -- parafermi supersymmetry and parasupersymmetry. The oscillator considered in this article also serves as a useful example to demonstrate the practical importance of the parabracket introduced in \cite{p10}. Moreover, it is remarkable to check that indeed all the conserved charges depend on the physical dynamical variables and not on their Green components. This is quite non-trivial, for all the calculations are performed using the Green components. In view of these observations, one may conclude that there is no anomalous phenomena stemming from the unusual parastatistical nature of the ($p=2$) dynamical variables. In fact, it is shown that for example the $\pi$SUSY oscillator has a larger symmetry than the ordinary SUSY oscillator. Another interesting observation regarding the symmetries of the $\pi$SUSY oscillator is that {\em a priori} there is no parity associated with the quantities (polynomials) constructed out of the parastatistical variables, nevertheless the conserved charges and the Hamiltonian do possess parities, and they do form a super Lie algebra. This may be seen as the primary reason why one does not need trilinear algebraic relations between the symmetry generators. The latter has been shown \cite{r-s} to be unavoidable for an oscillator consisting of ordinary bosons and ($p=2$) parafermions. The super Lie algebra associated with the $\pi$SUSY oscillator may be considered in a more general context. This line of reasoning leads to the introduction of supersymmetric paraquantum mechanics. The defining superalgebra of SPQM determines the degeneracy structure of the energy spectrum. The matrix reperesentation of the conserved charges reveals the differences and the similarities between SPQM and SQM. The Witten index can also be defined for SPQM. It possesses the topological invariance property and signifies the breaking of $\pi$SUSY, similarly to the ordinary SQM case. The Hilbert space structure of the $\pi$SUSY oscillator is also analyzed in detail. A remarkable observation is that the presence of ($p=2$) parafermi operators does not lead to a triple grading of the spectrum degeneracy. In fact, the general results obtained in the context of supersymmetric paraquantum mechanics are shown to be valid for the oscillator case. This serves as an independent check on the results obtained in Sec.~6. \section*{Acknowledgments} I would like to thank Kamran Saririan, Bahman Darian, Ertu\"grul Demircan and Stathis Tompaidis for mailing me some of the references. I wish to also acknowledge Shahin Rouhani for his constructive comments.
\section*{Figure Captions\markboth {FIGURECAPTIONS}{FIGURECAPTIONS}}\list {Figure \arabic{enumi}:\hfill}{\settowidth\labelwidth{Figure 999:} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endfigcap\endlist \relax \def\tablecap{\section*{Table Captions\markboth {TABLECAPTIONS}{TABLECAPTIONS}}\list {Table \arabic{enumi}:\hfill}{\settowidth\labelwidth{Table 999:} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endtablecap\endlist \relax \def\reflist{\section*{References\markboth {REFLIST}{REFLIST}}\list {[\arabic{enumi}]\hfill}{\settowidth\labelwidth{[999]} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endreflist\endlist \relax \def\list{}{\rightmargin\leftmargin}\item[]{\list{}{\rightmargin\leftmargin}\item[]} \let\endquote=\endlist \makeatletter \newcounter{pubctr} \def\@ifnextchar[{\@publist}{\@@publist}{\@ifnextchar[{\@publist}{\@@publist}} \def\@publist[#1]{\list {[\arabic{pubctr}]\hfill}{\settowidth\labelwidth{[999]} \leftmargin\labelwidth \advance\leftmargin\labelsep \@nmbrlisttrue\def\@listctr{pubctr} \setcounter{pubctr}{#1}\addtocounter{pubctr}{-1}}} \def\@@publist{\list {[\arabic{pubctr}]\hfill}{\settowidth\labelwidth{[999]} \leftmargin\labelwidth \advance\leftmargin\labelsep \@nmbrlisttrue\def\@listctr{pubctr}}} \let\endpublist\endlist \relax \makeatother \newskip\humongous \humongous=0pt plus 1000pt minus 1000pt \def\mathsurround=0pt{\mathsurround=0pt} \def\eqalign#1{\,\vcenter{\openup1\jot \mathsurround=0pt \ialign{\strut \hfil$\displaystyle{##}$&$ \displaystyle{{}##}$\hfil\crcr#1\crcr}}\,} \newif\ifdtup \def\panorama{\global\dtuptrue \openup1\jot \mathsurround=0pt \everycr{\noalign{\ifdtup \global\dtupfalse \vskip-\lineskiplimit \vskip\normallineskiplimit \else \penalty\interdisplaylinepenalty \fi}}} \def\eqalignno#1{\panorama \tabskip=\humongous \halign to\displaywidth{\hfil$\displaystyle{##}$ \tabskip=0pt&$\displaystyle{{}##}$\hfil \tabskip=\humongous&\llap{$##$}\tabskip=0pt \crcr#1\crcr}} \relax \hybrid \def\sigma{\sigma} \def\fnsymbol{footnote}{\fnsymbol{footnote}} \def\begin{equation}{\begin{equation}} \def\end{equation}{\end{equation}} \def\begin{eqnarray}{\begin{eqnarray}} \def\end{eqnarray}{\end{eqnarray}} \def\alpha{\alpha} \def\beta{\beta} \def\gamma{\gamma} \def\partial{\partial} \def\bar \partial{\bar \partial} \def\vartheta{\vartheta} \def\bar{J}{\bar{J}} \def\bar{\vartheta}{\bar{\vartheta}} \def\Phi{\Phi} \def\rho{\rho} \def\lambda{\lambda} \def\hbox{{$\sqcup$}\llap{$\sqcap$}}{\hbox{{$\sqcup$}\llap{$\sqcap$}}} \def\tau{\tau} \def{\rm Im}\tau{{\rm Im}\tau} \def{\cal{R}}{{\cal{R}}} \def{\cal{Q}}{{\cal{Q}}} \def{\rm I\!R}{{\rm I\!R}} \def\nabla{\nabla} \def\bar {\cal P}{\bar {\cal P}} \def{\cal R}{{\cal R}} \font\fivesans=cmss10 at 4.61pt \font\sevensans=cmss10 at 6.81pt \font\tensans=cmss10 \newfam\sansfam \textfont\sansfam=\tensans\scriptfont\sansfam= \sevensans\scriptscriptfont \sansfam=\fivesans \def\fam\sansfam\tensans{\fam\sansfam\tensans} \def{\partial\hspace{-.22cm}/}{{\partial\hspace{-.22cm}/}} \def\Z{{\mathchoice {\hbox{$\fam\sansfam\tensans\textstyle Z\kern-0.4em Z$}} {\hbox{$\fam\sansfam\tensans\textstyle Z\kern-0.4em Z$}} {\hbox{$\fam\sansfam\tensans\scriptstyle Z\kern-0.3em Z$}} {\hbox{$\fam\sansfam\tensans\scriptscriptstyle Z\kern-0.2em Z$}}}} \newlength{\boxsize} \def\limit#1#2{ \smash{ \mathop{#1} \limits_{#2} } } \def\limitt#1#2#3{ \settowidth{\boxsize}{\scriptsize $#2$} \, \makebox{\makebox[\boxsize][c]{$#1$} \hspace{-\boxsize} \hspace{-12pt} \raisebox{-4pt}{\scriptsize $#2$} \hspace{-\boxsize} \hspace{-12pt} \raisebox{-9pt}{\scriptsize $#3$} } } \def\stackrel{\leftrightarrow}{\partial}{\stackrel{\leftrightarrow}{\partial}} \def\stackrel{\leftrightarrow}{\bar\partial}{\stackrel{\leftrightarrow}{\bar\partial}} \begin{document} \renewcommand{\thesection.\arabic{equation}}}{\thesection.\arabic{equation}} \newcommand{\begin{equation}}{\begin{equation}} \newcommand{\eeq}[1]{\label{#1}\end{equation}} \newcommand{\begin{eqnarray}}{\begin{eqnarray}} \newcommand{\eer}[1]{\label{#1}\end{eqnarray}} \begin{titlepage} \begin{center} \hfill CERN-TH/95-171\\ \hfill LPTENS-95/28\\ \hfill hep-th/9508078\\ \vskip .2in {\large \bf Infrared behavior of Closed Superstrings in Strong Magnetic and Gravitational Fields} \vskip .4in {\bf Elias Kiritsis and Costas Kounnas\footnote{On leave from Ecole Normale Sup\'erieure, 24 rue Lhomond, F-75231, Paris, Cedex 05, FRANCE.}}\\ \vskip .3in {\em Theory Division, CERN,\\ CH-1211, Geneva 23, SWITZERLAND} \footnote{e-mail addresses: KIRITSIS,<EMAIL>}\\ \vskip .3in \end{center} \vskip .2in \begin{center} {\bf ABSTRACT } \end{center} \begin{quotation}\noindent A large class of four-dimensional supersymmetric ground states of closed superstrings with a non-zero mass gap are constructed. For such ground states we turn on chromo-magnetic fields as well as curvature. The exact spectrum as function of the chromo-magnetic fields and curvature is derived. We examine the behavior of the spectrum, and find that there is a maximal value for the magnetic field $H_{\rm max}\sim M_{\rm planck}^2$. At this value all states that couple to the magnetic field become infinitely massive and decouple. We also find tachyonic instabilities for strong background fields of the order ${\cal O}(\mu M_{\rm planck})$ where $\mu$ is the mass gap of the theory. Unlike the field theory case, we find that such ground states become stable again for magnetic fields of the order ${\cal O}(M^2_{\rm planck})$. The implications of these results are discussed. \end{quotation} \vskip 2.cm CERN-TH/95-171 \\ August 1995\\ \end{titlepage} \vfill \eject \def1.2{1.2} \baselineskip 16 pt \noindent \section{Introduction} \setcounter{equation}{0} In four-dimensional Heterotic or type II Superstrings it is possible in principle to understand the response of the theory to non-zero gauge or gravitational field backgrounds including quantum corrections. This problem is difficult in its full generality since we are working in a first quantized framework. In certain special cases, however, there is an underlying 2-d superconformal theory which is well understood and which describes exactly (via marginal deformations) the turning-on of non-trivial gauge and gravitational backgrounds. This exact description goes beyond the linearized approximation. In such cases, the spectrum can be calculated exactly, and it can provide interesting clues about the physics of the theory. In field theory (excluding gravity) the energy shifts of a state due to the magnetic field have been investigated long ago \cite{field,qcd,stand}. The classical field theory formula for the energy of a state of a spin $S$, mass $M$ and charge $e$ in a magnetic field $H$ pointing in the third direction is: \begin{equation} E^2=p_3^2+M^2+|eH|(2n+1-gS)\label{class} \end{equation} were $g=1/S$ for minimally coupled states and $n=1,2,...$ labels the Landau levels. It is obvious from (\ref{class}) that minimally coupled particles cannot become tachyonic, so the theory is stable. For non-minimally coupled particles, however, the factor $2n+1-gS$ can become negative and instabilities thus appear. For example, in non-abelian gauge theories, there are particles which are not minimally coupled. In the standard model, the $W^{\pm}$ bosons have $g=2$ and $S=\pm 1$. From (\ref{class}) we obtain that the spontaneously broken phase in the standard model is thus unstable for magnetic fields that satisfy \cite{qcd,stand} \begin{equation} |H|\geq {M_{W}^2\over |e|} \label{sta} \end{equation} A phase transition has to occur by a condensation of $W$ bosons, most probably to a phase where the magnetic field is confined (localized) in a tube, \cite{stand}. This behavior should be contrasted to the constant electric field case where there is particle production \cite{sch} for any non-zero value of the electric field, but the vacuum is stable (although the particle emission tends to decrease the electric field). The instabilities present for constant magnetic fields are still present in general for slowly varying (long range) magnetic fields. For a non-abelian gauge theory in the unbroken phase, since the mass gap is classically zero, we deduce from (\ref{class}) that the trivial vacuum ($A_{\mu}^{a}=0$) is unstable even for infinitesimally small chromo-magnetic fields. This provides already an indication at the classical level that the trivial vacuum is not the correct vacuum in an unbroken non-abelian gauge theory. We know however, that such a theory acquires a non-perturbative mass gap, $\Lambda^2\sim \mu^2 \exp[-16\pi^2 b_{0}/g^2]$ where $g$ is the non-abelian gauge coupling. If in such a theory one managed to create a chromo-magnetic field then there would again appear an instability and the theory would again confine the field in a flux tube. In string theory, non-minimal gauge couplings are present not only in the massless sector but also in the massive (stringy) sectors \cite{fpt}. Thus one would expect similar instabilities. Since in string theory there are states with arbitrary large values of spin and one can naively expect that if $g$ does not decrease fast enough with the spin (as is the case in open strings where $g=2$ \cite{fpt}) then for states with large spin an arbitrarily small magnetic field would destabilize the theory. This behavior would imply that the trivial vacuum is unstable. This does not happen however since the masses of particles with spin also become large when the spin gets large. The spectrum of open bosonic strings in constant magnetic fields was derived in \cite{acny}. Open bosonic strings however, contain tachyons even in the absence of background fields. It is thus more interesting to investigate open superstrings which are tachyon-free. This was done in \cite{fp}. It was found that for weak magnetic fields the field theory formula (\ref{class}) is obtained, and there are similar instabilities. In closed superstring theory however, one is forced to include the effects of gravity. A constant magnetic field for example carries energy, thus, the space cannot remain flat anymore. The interesting question in this context is, to what extend, the gravitational backreaction changes the behavior seen in field theory and open string theory. Such questions can have potential interesting applications in string cosmology since long range magnetic fields can be produced at early stages in the history of the universe where field theoretic behavior can be quite different from the stringy one. The first example of an exact electromagnetic solution to closed string theory was described in \cite{bas}. The solution included both an electric and magnetic field (corresponding to the electrovac solution of supergravity). In \cite{bk} another exact closed string solution was presented (among others) which corresponded to a Dirac monopole over $S^2$. More recently, several other magnetic solutions were presented corresponding to localized \cite{melvin} or covariantly constant magnetic fields \cite{rt}. The spectrum of these magnetic solutions seems to have a different behavior as a function of the magnetic field, compared to the situation treated in this paper. The reason for this is that \cite{rt} considered magnetic solutions where the gravitational backreaction produces a non-static metric. ``Internal" magnetic fields of the type described in \cite{bk} were also considered recently \cite{b} in order to break spacetime supersymmetry. Here we will study the effects of covariantly constant (chromo)magnetic fields, $H^a_i=\epsilon^{ijk}F_{jk}^a$ and constant curvature ${\cal R}^{il}=\epsilon^{ijk}\epsilon^{lmn}{\cal R}_{jm,kn}$, in four-dimensional closed superstrings. The relevant framework was developed in \cite{kk} where ground states were found, with a continuous (almost constant) magnetic field in a weakly curved space. We will describe in this paper the detailed construction of such ground states and we will eventually study their behavior. In the heterotic string (where the left moving sector is N=1 supersymmetric) the part of the $\sigma$-model action which corresponds to a gauge field background $A^{a}_{\mu}(x)$ is \begin{equation} V=(A^{a}_{\mu}(x)\partial x^{\mu}+F^{a}_{ij}(x)\psi^{i}\psi^{j})\bar J^{a} \label{vertex} \end{equation} where $F^{a}_{ij}$ is the field strength of $A^{a}_{\mu}$ with tangent space indices, eg. $F^{a}_{ij}=e^{\mu}_{i}e^{\nu}_{j}F^{a}_{\mu\nu}$ with $e^{\mu}_{i}$ being the inverse vielbein, and $\psi^{i}$ are left-moving world-sheet fermions with a normalized kinetic term. $\bar J^{a}$ is a right moving affine current. Consider a string ground state with a flat non-compact (euclidean) spacetime (${\rm I\!R}^{4}$). The simplest case to consider is that of a constant magnetic field, $H^a_i=\epsilon^{ijk}F_{jk}^a$. Then the relevant vertex operator (\ref{vertex}) becomes \begin{equation} V_{flat}=F_{ij}^a({1\over 2}x^{i} \partial x^{j}+\psi^{i}\psi^{j})\bar J^{a} \label{vertexflat} \end{equation} This vertex operator however, cannot be used to turn on the magnetic field since it is not marginal (when $F_{ij}^{a}$ is constant). In other words, a constant magnetic in flat space does not satisfy the string equations of motion, in particular the ones associated with the gravitational sector. A way to bypass this problem we need to switch on more background fields. In \cite{kk} we achieved this in two steps. First, we found an exact string ground state in which ${\rm I\!R}^{4}$ is replaced by ${\rm I\!R}\times S^{3}$. The ${\rm I\!R}$ part corresponds to free boson with background charge $Q=1/\sqrt{k+2}$ while the $S^{3}$ part corresponds to an $SU(2)_{k}$ WZW model. For any (positive integer) k, the combined central charge is equal to that of ${\rm I\!R}^{4}$. For large $k$, this background has a linear dilaton in the $x^{0}$ direction as well as an $SO(3)$-symmetric antisymmetric tensor on $S^{3}$, while the metric is the standard round metric on $S^{3}$ with constant curvature. On this space, there is an exactly marginal vertex operator for a magnetic field which is \begin{equation} V_{m}=H(J^{3}+\psi^{1}\psi^{2})\bar J^{a}\label{magnet} \end{equation} Here, $J^{3}$ is the left-moving current of the $SU(2)_{k}$ WZW model. $V_{m}$ contains the only linear combination of $J^{3}$ and $\psi^{1}\psi^{2}$ that does not break the N=1 local supersymmetry. The exact marginality of this vertex operator is obvious since it is a product of a left times a right abelian current. This operator is unique up to an $SU(2)_{L}$ rotation. We can observe that this vertex operator provides a well defined analog of $V_{flat}$ in eq. (\ref{vertex}) by looking at the large $k$ limit. We will write the SU(2) group element as $g=\exp[i{\vec\sigma}\cdot{\vec x}/2]$ in which case $J^{i}=kTr[\sigma^{i} g^{-1}\partial g]=ik(\partial x^{i}+\epsilon^{ijk}x_j\partial x_k+{\cal{O}}(|x|^3))$. In the flat limit the first term corresponds to a constant gauge field and thus pure gauge so the only relevant term is the second one which corresponds to constant magnetic field in flat space. The fact $\pi_{2}(S^{3})=0$ explains in a different way why there is no quantization condition on $H$. THis magnetic field background break spacetime supersymmetry as usually expected. It evades one of the assumptions of the Banks-Dixon theorem \cite{BD} since it involves a vertex operator from non-compact 4-d spacetime. There is another exactly marginal perturbation in the background above that turns on fields in the gravitational sector. The relevant perturbation is \begin{equation} V_{grav}={\cal{R}} (J^{3}+\psi^{1}\psi^{2})\bar J^{3} \label{gravi}\end{equation} This perturbation modifies the metric, antisymmetric tensor and dilaton \cite{kk}. For type II strings the relevant perturbation is \begin{equation} V_{grav}^{II}={\cal{R}}(J^3+\psi^{1}\psi^{2})(\bar J^{3}+\bar\psi^{1}\bar \psi^{2}) \end{equation} The space we are using, ${\rm I\!R}\times S^3$ is such that the spectrum has a mass gap $\mu^2$. In particular all gauge symmetries are broken spontaneously. This breaking however is not the standard breaking due to a constant expectation value of a scalar but due to non-trivial expectation values of the fields in the universal sector (graviton, antisymmetric tensor and dilaton). In subsequent sections we derive the exact spectrum of closed string ground states in the presence of magnetic and gravitational fields generated by (\ref{magnet}), (\ref{gravi}). An interesting first result is that such magnetic fields in closed strings cannot become larger than a maximal value \begin{equation} H_{\rm max}={M^2_{\rm planck}\over \sqrt{2}} \end{equation} where $M_{\rm planck}=M_{\rm string}/g_{\rm string}$, $M_{\rm string}=1/\sqrt{\alpha'}$ and $g_{\rm string}$ is the string coupling constant. This is reminiscent of the appearance of a (finite) maximal electric field in open superstrings \cite{bp}. There, the maximal electric field is associated with the value at which the pair-production rate per unit volume becomes infinite. Here at $H=H_{\rm max}$ all states that couple to the magnetic field (i.e. having non-zero charge and/or angular momentum) become infinitely massive. This phenomenon is similar to the limit ${\rm Im} U\to \infty$ in a 2-d torroidal CFT. It is thus a boundary of the magnetic field moduli space. Since the theories we consider have a mass gap, they are stable for small magnetic fields. As we keep increasing the magnetic field we find an instability for $|H|\geq H^{\rm crit}_{\rm lower}$ whose origin is similar to the field theoretic one, namely states with helicity one and non-minimal coupling become tachyonic. Unlike field theory though (where $H^{\rm crit}_{\rm lower}\sim \mu^2$) here we find \begin{equation} H^{\rm crit}_{\rm lower}\sim \mu M_{\rm planck} \end{equation} The reason for the different behavior can be traced to the way the gauge symmetry is broken in field theory versus our string ground states. In field theory higher helicity particles with non-minimal couplings (like charged gauge bosons) get a mass from the expectation value of a charged scalar (Higgs field). In the string ground states we consider, the gauge symmetries are broken by non-trivial expectations values in the generic sector (graviton, antisymmetric tensor, dilaton). Another major difference with field theory behavior is the following: In field theory, for $H\geq H^{\rm crit}_{\rm lower}$ the theory is unstable for arbitrarily high magnetic fields. In closed string theory we find that theory generically becomes stable again for \begin{equation} H^{\rm crit}_{\rm upper}\leq |H|\leq H_{\rm max}\label{up} \end{equation} with \begin{equation} H_{\rm max}-H^{\rm crit}_{\rm upper}\sim \mu M_{\rm planck} \end{equation} One might think that this region is irrelevant since the theory undergoes a phase transition already for smaller magnetic fields. One however, could imagine a situation where a strong localized magnetic field (which does not induce an instability) starts spreading out in space to become a long range field in the region described by (\ref{up}). The geometry of spacetime, although deformed by the presence of the magnetic field, remains smooth (free of singularities) for the whole range $0\leq |H| \leq H_{\rm max}$. Similar remarks apply to the gravitational perturbation (\ref{gravi}). There, we can describe this perturbation with a modulus $0\leq \lambda\leq 1$ so that $\lambda=1$ corresponds to the round three-sphere geometry. For $\lambda\not= 1$ the sphere is squashed as can be seen from the expression of the scalar curvature (in Euler angles): \begin{equation} R_{\rm scalar} ={8\over k}{-1+5\lambda^2-\lambda^4+2H^2\lambda^2+(1-\lambda^4)\cos\beta\over ( 1+\lambda^2+(\lambda^2-1)\cos\beta)^2} \end{equation} where we have also included the effects of a magnetic field $H$. We note that again the geometry is smooth for $\lambda\not= 0$, while it becomes singular at $\lambda=0$. This singularity corresponds to the classical singularity of the $SU(2)_{k}/U(1)$ coset model \cite{gk} which is known to be absent from the corresponding CFT. Simply, at $\lambda=0$ one dimension decompactifies. Here again we find an instability for \begin{equation} 0\leq \lambda_{\rm lower}\leq \lambda \leq \lambda_{\rm upper}\leq 1 \end{equation} The structure of this paper is as follows. In section 2 we describe how to construct, for any given 4-d flat space ground state of the superstring, another ground state in curved 4-d space, with similar matter spectrum, but with a non-zero mass gap $\mu^2$. In section 3 we give the $\sigma$-model description of turning on non-trivial magnetic fields and curvature in the ground states described in section 2. In section 4 we give the CFT description of such magnetic fields. In section 5 we discuss the flat space limit when the mass gap $\mu^2\to 0$. Finally section 6 contain our conclusions and further directions. \section{Construction of curved ${W_{k}}\otimes K$ string solutions} \setcounter{equation}{0} Our aim is to replace the Euclidean four-dimensional flat space solution ${\rm I\!R}^4\otimes K$ by a curved space solution $W_{k}\otimes K$, where we replace the four non-compact (super)coordinates of flat space by the (super)coordinates of the $SU(2)_k\otimes {\rm I\!R}_Q$ theory. Three of the coordinates describe $SU(2)_k$ (the three-dimensional sphere) and the fourth is a flat coordinate with non zero background charge $Q=1/\sqrt{k+2}$. The relation among the level $k$ (a non-negative integer) and the background charge $Q$ is such that the left and right central charges remains the same as in ${\rm I\!R}^4$ for any value of $k$. This give us the possibility to keep unchanged the internal superconformal theory $K$. We will show below that the replacement ${\rm I\!R}^4\rightarrow W_k$ can be done in a universal way and without reducing the number of spacetime supersymmetries in almost all interesting cases with non-maximal number of supersymmetries. In the case of maximal supersymmetry, ($N=4$ in heterotic and $N=8$ in type II), this replacement, although is still universal, reduces by a factor of two the number of space time supersymmetries. This reduction is unavoidable and due to the fact that only half of the constant killing spinors of ${\rm I\!R}^4$ remain covariantly constant in the $W_k$ space \cite{wormclas,worm,bil}. The non-zero torsion and dilaton are responsible for this. Let us start first with the case of maximal supersymmetry in flat space. Following \cite{BD}, the world sheet super-current(s) of the internal $K$ theory can be always constructed in terms of six free bosons compactified on a torus and six free fermions\footnote{There are some exotic cases though \cite{cp}, which will not be discussed here.}; the internal fermions, the ${\rm I\!R}^4$ fermions and the $\beta, \gamma$ ghosts of supereparametrizations must have identical boundary conditions (periodic or antiperiodic) in all non-trivial worldsheet cycles. In type II, this global restriction must be respected separately for the left- and right-moving fermions and the $\beta, \gamma$ ghosts. Also, the left and right momenta for the compactified bosons in $K$ must form necessarily a self-dual lorentzian lattice. Both ${\rm I\!R}^4$ and $W_k$ are $N=4$ ${\hat c}=4$ superconformal theories. In both cases the $SU(2)_1$ $N=4$ currents $S^i$, $i=1,2,3$ are constructed in terms of word sheet fermions $\psi_{0}$, $\psi_i$, $i=1,2,3$ \begin{equation} S^{i}= \frac{1}{2}\left(\psi_{0}\psi^{i}+ \frac{1}{2}\epsilon^{ijl}\psi_{j}\psi_{l}\right)\ . \end{equation} Observe that only the three self-dual currents appear in the algebra. It order to specify better the difference among the ${\rm I\!R}^4$ and $W_k$ theories, it is convenient to parametrize the $S^{i}$ (self-dual) currents and the remaining anti self-dual ones ${\tilde S}^i$ \begin{equation} {\tilde S}^{i} = \frac{1}{2}\left(-\psi_{0}\psi^{i}+ \frac{1}{2}\epsilon^{ijl}\psi_{j}\psi_{l}\right)\ . \end{equation} in terms of two free bosons, $H^{+}$ and $H^{-}$, both compactified on a circle with radius $R_{H^{+}}=R_{H^{-}}=1$ (the self-dual $SU(2)$ extended symmetry point). In both cases, the four $N=4$ supercurrents $G$, $G^{\dag}$, $\bar G$ and $\bar G^{\dag}$ are given as \cite{k} \begin{equation} {G} = -\left( \Pi^{\dag}e^{-\frac{i}{\sqrt{2}}H^{-}} + P^{\dag}e^{+ \frac{i}{\sqrt{2}} H^{-}} \right) e^{+ \frac{i}{ \sqrt{2}}H^{+}} \label{compl1}\end{equation} \begin{equation} {\bar {G}} = \left( \Pi \;\; e^{+\frac{i}{\sqrt{2}}H^{-}} - P\;\; e^{-\frac{i}{ \sqrt{2}} H^{-}} \right) e^{+ \frac{i}{\sqrt{2}}H^{+}} \label{compl2} \end{equation} where $P$, $P^{\dag}$ and $\Pi$, $\Pi^{\dag}$ are the four coordinate currents. In ${\rm I\!R}^4$ they are \begin{equation} \Pi = \partial X_0+i\partial X_3 \, , \;\;\;\;\; \Pi^{\dag}= -\partial X_0 +i\partial X_3 \, \end{equation} \begin{equation} P= \partial X_1 +i\partial X_2 \, , \;\;\;\;\; P^{\dag}= -\partial X_1 +i\partial X_2 \, \label{complbF} \end{equation} In the $W_k$ case, $P$, $P^{\dag}$ and $\Pi$, $\Pi^{\dag}$ get modified due to the torsion and non-trivial dilaton. They can be constructed in terms of the $SU(2)_k\otimes {\rm I\!R}_Q$ (anti-hermitian) currents $J_i$ $i=1,2,3$, $J_0=\partial x^{0}$ and $H^{-}$, $$ \Pi=J_0 + iQ(J_3+\sqrt{2}\partial H^-), \,\,\,\,\, \Pi^{\dag}=-J_0+ iQ(J_3+\sqrt{2}\partial H^-)\ , $$ \begin{equation} P=Q(J_1+iJ_2), \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\ P^{\dag}=Q(-J_1+iJ_2)~~~~~~~~~~~~~~ \label{Pcompl} \end{equation} The $H^-$ modifications are due to the non-trivial background fields of the $W_{k}$ space. In terms of the bosonized fermions, both the standard fermionic torsion terms $\pm Q\psi_i\psi_j\psi_l$ and the fermionic background charge terms ($\pm Q\partial\psi_i$) are combined in such a way that only the anti-self dual combination, $\partial H^-$, modifies the coordinate currents. The other important observation is the $H^{+}$ part in the supercurrents is factorized. The $H^{+}$ factorization property is universal in all $N=4$ superconformal ${\hat{c}}=4$ superconformal theories \cite{k}. It is evident from (\ref{compl1}), (\ref{compl2}) and (\ref{Pcompl}) that the supercurrents (${G},G^{\dag}$) and (${\bar {G}},{\bar{G}}^{\dag}$) form two doublets under $SU(2)_{H^{+}}$. On the other hand, ${G}$ and ${\tilde {G}}$ do not transform covariantly under the action of $SU(2)_{H^{-}}$. They are odd, however, under a $\Z_2$ transformation, defined by $(-)^{2{\tilde S}}$, which is the parity operator associated to the $SU(2)_{H^{-}}$ spin ${\tilde S}$ (integer spin representations are even, while half-integer representations are odd). In $W_k$ we can define a global $SU(2)_{k+1}$ charge as the diagonal combination of $SU(2)_k$ and $SU(2)_{H^-}$: \begin{equation} {\cal N}_i = J_i + {\tilde S}_i\ . \label{ninumb} \end{equation} (${G},G^{\dag}$) and (${\bar {G}},{\bar {G}}^{\dag}$) form two doublets under this $SU(2)_{k+1}$. Moreover $G$ and ${\bar {G}}$ have $({\cal N}_3, S_3)$ charges equal to $(-1/2,1/2)$ and $(1/2,1/2)$, respectively. The global charge ${\cal N}_3$ in $W_k$ plays the role of the helicity operator $N_h$ of flat space \begin{equation} N_h=N_p+{\tilde S}_3 \end{equation} where $N_p$ is the bosonic oscillator number which counts the number of the $P$-oscillators minus the number of $P^{\dag}$ ones. The $N_h$, ${\cal N}_3$ charge, the $(-)^{2{\tilde S}}$ parity, as well as the $SU(2)_{H^{+}}$ spin $S$ play an important role in the definition of the induced generalized GSO projections of the unitary $N=4$ characters. We would now like to show that when flat Euclidean four-space is replaced by $W_k$, the maximal supersymmetry is reduced by a factor of two. In order to avoid heavy notation for the vertex operators which include the reparametrization ghosts, it is convenient to start our discussion with a six dimensional theory ${\rm I\!R}^2\otimes {\rm I\!R}^4\otimes K$ and compare it with ${\rm I\!R}^2\otimes W_k\otimes K$. In the type-II case, both the flat and curved constructions have their degrees of freedom arranged in three superconformal theories as: \begin{equation} \{ \hat c\} = 10 = \{ \hat c = 2 \} + \{ \hat c = 4 \}_1 + \{ \hat c = 4\}_2~. \label{ceq} \end{equation} The $\hat c = 2$ system is saturated by two free superfields. The remaining eight supercoordinates appear in groups of four in $\{\hat c = 4\}_1$ and $\{\hat c = 4\}_2$. Both $\{\hat c = 4\}_A$ systems exhibit $N = 4$ superconformal symmetry of the Ademollo et al. type \cite{ademolo}. The advantage of the six dimensional space in which two super-coordinates are flat is that we can use the light-cone picture in which the two-dimensional subspace ${\rm I\!R}^2$ is flat (non-compact with Lorentzian signature) and the eight transverse coordinates are described by the $\{\hat c = 4\}_1$ and $\{\hat c = 4\}_2$ theories. In this picture, the supersymmetry generators are constructed by analytic (or antianalytic) dimension-one currents, whose transverse part is a spin-field of dimension 1/2, constructed in terms of the $H^+_A$ and $H^-_A$ bosonized fermions (of the two ${\hat c}_A=4$, $A=1,2$ theories). In the toroidal case, there are four such spin-fields, \begin{eqnarray} \Theta_{\pm} &=& e^{\frac{i}{\sqrt 2}(H^+_1 \pm H^+_2)}\nonumber \\ {\rm and}~~~~\tilde{\Theta}_{\pm} &=& e^{\frac{i}{\sqrt 2}(H^-_1 \pm H^-_2)}\ . \label{susy} \end{eqnarray} which are even under the GSO parity: \begin{equation} e^{2i\pi (S_1 +S_2)}\ , \label{GSO} \end{equation} where $S_A$ are the two $SU(2)_{H^+_A}$ level-one $N=4$ spins. In the case of ${\rm I\!R}^2\otimes W_k\otimes K$ , only the two supersymmetry generators based on the operators $\Theta_{\pm}$, which are constructed with $H^+$'s, are BRS-invariant. Indeed, the other two operators $\tilde{\Theta}_{\pm}$ are not physical, due to the presence of the $\partial H^-$ modification, related to the torsion and/or background charge, in the supercurrent expressions. The global existence of the (chiral) $N=4$ superconformal algebra implies in both constructions, a universal GSO projection that generalizes \cite{worm} the one of the $N=2$ algebra \cite{lls,gepner4d}, and which is responsible for the existence of space-time supersymmetry. This projection restricts the physical spectrum to being odd under the total $H^+_A$ parity (\ref{GSO}). Thus, the supersymmetry generators based on $\Theta_{\pm}$, which are even under (\ref{GSO}), when acting on physical states, create physical states with the same mass but with different statistics. The GSO projection restricts the (level-one) character combinations associated with the two $SU(2)_{H^+}$'s to appear in the form: \begin{equation} \frac{1}{2} (1-(-)^{l_1+l_2}) \chi^{H^+_1}_{l_1}\chi^{H^+_2}_{l_2} = \chi^{H^+_1}_{l_1} \chi^{H^+_2}_{1-l_1}\delta_{l_2,1-l_1}\ , \label{charplus} \end{equation} with $l_A=2S_A$ taking values 0 or 1, corresponding to the two possible characters, (spin-$0$ and spin-$1/2$) of the $SU(2)_{1}$ affine algebra. The basic rules in both constructions are similar to those of the orbifold construction \cite{orbifold}, the free 2-d fermionic constructions \cite{ferm}, and the Gepner construction \cite{gepner4d}. There, one combines in a modular invariant way the world-sheet degrees of freedom consistently with unitarity and spin-statistics of the string spectrum. In both cases, the 2-d fermions are free and their characters are given in terms of $\vartheta$-functions. The 6-d Lorentz invariance in the flat case and the existence of $SU(2)_k$ currents in the curved case imply the same boundary conditions for the super-reparametrization ghosts and the six of the worldsheet fermions. In the flat case there is no obstruction to choose the remaining four fermions with the same boundary conditions and obtain the well known partition function with maximal space-time supersymmetry: \begin{eqnarray} Z_F = {1\over {\rm Im}\tau^2|\eta|^8}~{1\over 4} \sum_{\alpha,\beta\atop {\bar\alpha},{\bar\beta}}~ (-)^{\alpha+\beta+{\bar\alpha}+{\bar\beta}} \frac{\vartheta^{2}[^{\alpha}_{\beta}]}{\eta^2} \frac{\vartheta^{2}[^{\alpha}_{\beta}]}{\eta^2}~ \frac{{\bar\vartheta}^{2}[^{\bar\alpha}_{\bar\beta}]}{{\bar\eta}^2} \frac{{\bar\vartheta}^{2} [^{{\bar\alpha}}_{{\bar\beta}}]}{{\bar\eta}^2}~Z_4[^0_0], \label{pF} \end{eqnarray} where the $Z_4[^0_0]$ contribution is that of four compactified coordinates: \begin{equation} Z_{4}[^0_0]={\Gamma(4,4)\over |\eta|^8}\label{444} \end{equation} and $\Gamma(4,4)$ stands for the usual lattice sum. $\alpha$, $\beta$ and ${\bar\alpha}$, ${\bar\beta}$ denote the left- and right-moving spin structures. The spin-statistic factors $(-)^{\alpha +\beta}$ and $(-)^{\bar{\alpha} +\bar{\beta}}$ come from the contribution of the (left- and right-moving) ${\rm I\!R}^2$ world-sheet fermions and the (left- and right-moving) $({\bf\beta}, {\bf\gamma})$-ghosts. The Neveu-Schwarz ($NS$, $\overline{NS}$) sectors correspond to $\alpha, \bar{\alpha}=0$ and the Ramond ($R,\bar{R}$) sectors correspond to $\alpha, \bar{\alpha}=1$. For later convenience we decompose the $O(4)$ level-one characters, which are written in terms of $\vartheta$-functions, in terms of the $SU(2)_{H^{+}_{1}}\otimes SU(2)_{H^{-}_{1}}$ characters using the identity: \begin{equation} \frac{\vartheta ^{2} [^{\alpha}_{\beta}]}{ \eta ^{2}(\tau)}= \sum_{l=0}^{1} (-)^{\beta l} \chi_{l}^{H^{+}_1} \chi_{l+\alpha(1-2l)}^{H^{-}_1}\ , \label{theeta} \end{equation} and similarly for the right-movers. Using the decomposition above, we can write the flat partition function in terms of $SU(2)_{H^{+}_1}$, $SU(2)_{H^{-}_2}$ , $SU(2)_{H^{+}_2}$ and $SU(2)_{ H^{-}_2 }$ characters as \begin{eqnarray} Z_F = {1\over {\rm Im}\tau^2|\eta|^8} {}~\sum_{\alpha,{\bar\alpha}}~ (-)^{\alpha+{\bar\alpha}}~ \chi_{l}^{H^{+}_1} \chi_{l+\alpha}^{H^{-}_1}~ \chi_{1+l}^{H^{+}_2} \chi_{1+l+\alpha}^{H^{-}_2}~ {\bar\chi}_{\bar l}^{H^{+}_1} {\bar\chi}_{{\bar l}+{\bar\alpha}}^{H^{-}_1}~ {\bar\chi}_{1+{\bar l}}^{H^{+}_2} {\bar\chi}_{1+{\bar l}+{\bar\alpha}}^{H^{-}_2} {}~Z_4[^0_0], \label{pFchar} \end{eqnarray} In going from (\ref{pF}) to (\ref{pFchar}), the $\beta$ and ${\bar\beta}$ summations give rise to the universal (left- and right-moving) GSO projections among the $SU(2)_{H^{+}_1}$ and $SU(2)_{H^{+}_2}$ spins, $2 S_1 + 2{ S}_2$=odd, as well as among those of $SU(2)_{H^{-}_1}$, $SU(2)_{H^{-}_2}$, $2\tilde S_{1}+2\tilde S_{2}=$ odd. These projections imply the existence of maximal space-time supersymmetry. The phase $(-)^{\alpha+{\bar\alpha}}$ guarantees the spin-statistics connection; it equals $+1$ for space-time bosons and $-1$ for space-time fermions. Replacing ${\rm I\!R}^4$ flat space with $W_k$, some modifications are necessary. Namely, we must combine the ${\rm I\!R}_Q$ Liouville-like characters and the $SU(2)_k$ ones ($\chi _{L}~,~L=0,1,2,\cdots, k$) with those of the remaining 2-d bosons and fermions, in a way consistent with unitarity and modular invariance. The ${\rm I\!R}_Q$ Liouville-like characters can be classified in two categories. Those that correspond to the continuous representations generated by the operators: \begin{equation} e^{\beta X_L}\ ;\quad \beta=-\frac{1}{2}Q +ip\ , \label{contchar} \end{equation} having positive conformal weights $\Delta_p=\frac{Q^2}{8}+\frac{p^2}{2}$. The fixed imaginary part in the momentum $iQ/2$ of the plane waves, is due to the non-trivial dilaton motion. The second category of Liouville characters \cite{sbk} corresponds to lowest-weight operators (\ref{contchar}) with $\beta=Q\tilde{\beta}$ real, leading to negative conformal dimensions $-\frac{1}{2}\tilde{\beta}(\tilde{\beta}+1)Q^2= -\frac{\tilde{\beta}(\tilde{\beta}+1)}{k+2}$. Both categories of Liouville representations give rise to unitary representations of the $N=4$, ${\hat c}=4$ $W_k$ theory, once they are combined with the remaining degrees of freedom. The continuous representations (\ref{contchar}) form long (massive) representations \cite{ademolo} of $N=4$ with conformal weights larger than the $N=4$, $SU(2)$ spin, $\Delta>S$. On the other hand, the second category contains short representations of $N=4$ \cite{ademolo} ($\Delta=S$), while $\beta$ can take only a finite number of values, $-(k+2)/2\le\tilde{\beta}\le k/2$. In fact, their locality with respect to the $N=4$ operators implies \cite{worm}: \begin{eqnarray} S &=& \frac{1}{2},\quad \tilde{S}=\frac{1}{2}:\quad \tilde{\beta} = -(j+1) \nonumber \\ S &=& 0,\quad \tilde{S}=0:\quad \tilde{\beta} = j\ . \label{discrchar} \end{eqnarray} In both cases of (\ref{discrchar}), the conformal weight $\Delta=S$ is independent of $SU(2)_k$ and $SU(2)_{H^-}$ spins, due to the cancellation between the Liouville and $SU(2)_k$ contributions. The states associated to the short representations of $N=4$ do not have the interpretation of propagating states, but they describe a discrete set of localized states. They are similar to the discrete states found in the $c=1$ matter system coupled to the Liouville field \cite{lz} and the two-dimensional $SL(2,R)/O(1,1)$ coset model \cite{sl2}. Although they play a crucial role in scattering amplitudes, they do not correspond to asymptotic states and they do not contribute to the partition function. Indeed, in our case they are not only discrete but also finite in number. They obviously are of zero measure compared to the contribution of the continuous (propagating) representations. The presence of discrete representations with $\beta$ positive are necessary to define correlation functions. In fact, the balance of the background charge for an $N$-point amplitude at genus $g$ implies the relation \cite{worm} \begin{equation} N+2(g-1)+2\sum_I\tilde{\beta}_I=0\ , \label{screen} \end{equation} where the sum is extended over the vertices of the discrete representation states. Thus, these vertices define an appropriate set of screening operators, necessary to define amplitudes in the presence of non-vanishing background charge. In our case, the screening procedure has an interesting physical interpretation similar to the scattering of asymptotic propagating states (continuous representations) in the presence of non-propagating bound states (discrete representations). The screening operation then describes the possible angular momentum (of SU(2)) excitations of the bound states. Below, we restrict ourselves to the one-loop partition function, where the discrete representations are not necessary (see eq.(\ref{screen}) with $g=1$ and $N=0$). It is convenient to define appropriate character combinations of $SU(2)_k$, which transform covariantly under modular transformations \cite{worm}: \begin{equation} Z_{so(3)}[^{\alpha}_{\beta}] =Z_{so(3)}[^{\alpha+2}_{\beta}]=Z_{so(3)}[^{\alpha}_{\beta+2}]= e^{-i\pi\alpha\beta k/2}\sum_{l=0}^k e^{i\pi\beta l} \chi_l \bar{\chi}_{(1-2\alpha)l+\alpha k} \label{zab} \end{equation} where $\alpha$, $\beta$ can be either 0 or 1. The $SU(2)_{k}$ characters are given by the familiar expressions \cite{kp}, \begin{equation} \chi_{l}(\tau)={\vartheta_{l+1,k+2}(\tau,v)-\vartheta_{-l-1,k+2}(\tau,v)\over \vartheta_{1,2}(\tau,v)-\vartheta_{-1,2}(\tau,v)}|_{v=0} \end{equation} where \begin{equation} \vartheta_{m,k}(\tau,v)\equiv \sum_{n\in \Z}\exp\left[2\pi i k\left(n+{m\over 2k}\right)^2\tau-2\pi i k\left(n+{m\over 2k}\right)v\right] \end{equation} are the level-k $\vartheta$-functions. The projection induced in (\ref{zab}) will project $SU(2)\to SO(3)$. This can be done consistently when $k$ is an even integer, which we assume from now on. Under modular transformations, $Z_{so(3)}[^{\alpha}_{\beta}]$ transforms as: \begin{eqnarray} \tau\rightarrow\tau +1~~~:~~~ &Z_{so(3)}[^{\alpha}_{\beta}]& \longrightarrow Z_{so(3)}[^{~\alpha}_{\beta+\alpha}] \nonumber \\ \tau\rightarrow{-1/\tau}~~~:~~~ &Z_{so(3)}[^{\alpha}_{\beta}]& \longrightarrow Z_{so(3)}[^{\beta}_{\alpha}]\ . \label{zabtransf} \end{eqnarray} The partition function must satisfy two basic constraints emerging from the $N=4$ algebra. The first is associated to the two spectral flows of the $N=4$ algebra which impose the universal GSO projection $2(S_1+S_2)$=odd among the $H^+$ spins. The second constraint is associated to the reduction of space-time supersymmetries by a factor of 2. It imposes a second projection which eliminates half of the lowest-lying states constructed from the $H^-_1$ field, which are not local with respect to the $N=4$ generators. These unphysical states should be eliminated from the spectrum by an additional GSO projection involving the two $H^{-}_i$ spins $\tilde S_i$ as well as the spin of $SU(2)_k$. For $k$ even, there is a $\Z_2$ automorphism of $SU(2)_k$ which leaves invariant the currents but acts non-trivially on the odd spin representations. This allows to correlate the $SU(2)_{H_1^-}$, $SU(2)_{H_1^-}$ and $SU(2)_k$ spins in a way which projects out of the spectrum the unphysical states. This $\Z_2$ must act simultaneously on the four toroidal compactified coordinates in order to guarantee the global existence of the $N=1$ supercurrent. The modular-invariant partition function then is: \begin{equation} Z_W = {1\over {\rm Im}\tau^{1/2}|\eta|^2}~{1\over 8} \sum_{\alpha,\beta,{\bar\alpha}\atop {\bar\beta},\gamma,\delta} (-)^{(\alpha+{\bar\alpha})(1+\delta)+\beta+{\bar\beta}} \frac{\vartheta^{2}[^{\alpha}_{\beta}]}{\eta^2} \frac{\vartheta^{2}[^{\alpha+\gamma}_{\beta+\delta}]}{\eta^2}~ \frac{{\bar\vartheta}^{2}[^{\bar\alpha}_{\bar\beta}]}{{\bar\eta}^2} \frac{{\bar\vartheta}^{2} [^{{\bar\alpha}+\gamma}_{{\bar\beta}+\delta}]}{{\bar\eta}^2} {Z_{so(3)}[^{\gamma}_{\delta}]\over V}~ Z_4[^\gamma _\delta] \label{pW} \end{equation} where $Z_4[^\gamma _\delta ]$ denotes the $T^{(4)}/\Z_2$ orbifold twisted characters. $Z_{4}[^0_0]$ is given in (\ref{444}) while for $(h,g)\not= (0,0)$ we have \begin{equation} Z_{4}[^h_g]={|\eta|^4\over |\vartheta[^{1+h}_{1+g}]\vartheta[^{1-h}_{1-g}]|^2} \end{equation} We have also divided by the (quantum) volume of $S^3$ \begin{equation} V={(k+2)^{3/2}\over 8\pi}\;\;\;.\label{volume} \end{equation} We may rewrite the partition function above in terms of various $SU(2)$ characters so that the induced GSO projections are more transparent. $$ Z_W = {1\over {\rm Im}\tau^{1/2}|\eta|^2} {}~\sum_{\alpha,{\bar\alpha},\gamma,l,{\bar l}=0}^{1} (-)^{\alpha+{\bar\alpha}}~ \chi_{l}^{H^{+}_1} \chi_{l+\alpha}^{H^{-}_1}~ \chi_{l+1}^{H^{+}_2} \chi_{l+1+\alpha+\gamma}^{H^{-}_2}~ {\bar\chi}_{\bar l}^{H^{+}_1} {\bar\chi}_{{\bar l}+{\bar\alpha}}^{H^{-}_1}~ {\bar\chi}_{1+{\bar l}}^{H^{+}_2} {\bar\chi}_{1+{\bar l}+{\bar\alpha}+\gamma}^{H^{-}_2}\times $$ \begin{equation} \times~{1\over V}\sum_{L=0}^{k}\sum_{\delta=0}^{1} {1\over 2} (-)^{\delta[\alpha+l+{\bar\alpha}+{\bar l}+{k\over 2}\gamma+L]}~ \chi_L \bar{\chi}_{(1-2\gamma)L+\gamma k}~Z_4[^{\gamma}_{\delta}]\ , \label{pWchar} \end{equation} As in the flat case, the $\beta$ and ${\bar\beta}$ summations give rise to the universal (left- and right-moving) GSO projections $2(S_1+S_2)$=odd, which imply the existence of space-time supersymmetry. The summation over $\delta$ however gives rise to an additional projection, which correlates the $SU(2)_{H^-_2}$ (left and right) spin together with the spin of $SU(2)_k$ and $T^{(4)}/\Z_2$ twisted bosonic oscillator numbers. This projection reduces the number of space-time supersymmetries by a factor of two. In the $\gamma=0$ sector (untwisted sector), the lower-lying states have (left and right) mass-squared $Q^2/8$ and $L=0$. This is due to the non-trivial dilaton for the bosons, and to the non-trivial torsion for the fermions. The contribution of 2-d fermions in the partition function of the $\gamma=0$ sector is identical to the fermionic part of the partition function of the ten-dimensional type II superstring with an additional projection acting on $SU(2)_k$ spins : \begin{equation} Z_W^{\gamma =0} = {1\over {\rm Im}\tau^{1/2}|\eta|^2} {1\over 4} |\vartheta_3^4-\vartheta_4^4-\vartheta_2^4|^2 \sum_{L=even}^{k}{ |\chi_L|^2\over V}~Z_4[^0_0] \label{pfu} \end{equation} As was stressed before, the extra $\Z_2$ projection is dictated from the $N=4$ superconformal algebra, in order to eliminate the unphysical states from the untwisted sector. Modular invariance implies the presence of a twisted sector $(\gamma=1)$, which contains states with (left and right) mass-squared always larger than $(k-2)/16$. In the large $k$ limit the twisted states become super-heavy We can now return to our initial problem and examine the Euclidean $W_k\otimes K$ theory with maximal space-time supersymmetry. The latter can be obtained by a $T^{2}$ torus compactification from the Euclidean version of the six dimensional construction described above. Observe that it is necessary to act non-trivially in the internal theory $K$, since the $\Z_2$ in question has to act on the four out of the six compactified coordinates. The resulting partition function is: \begin{equation} Z^{4d}_W = {{\rm Im}\tau^{1/2}|\eta|^2\over 8} \sum_{\alpha,\beta,{\bar\alpha}\atop {\bar\beta},\gamma,\delta} (-)^{(\alpha+{\bar\alpha})(1+\delta)+\beta+{\bar\beta}} \frac{\vartheta^{2}[^{\alpha}_{\beta}]}{\eta^2} \frac{\vartheta^{2}[^{\alpha+\gamma}_{\beta+\delta}]}{\eta^2} \frac{{\bar\vartheta}^{2}[^{\bar\alpha}_{\bar\beta}]}{{\bar\eta}^2} \frac{{\bar\vartheta}^{2} [^{{\bar\alpha}+\gamma}_{{\bar\beta}+\delta}]}{{\bar\eta}^2} {Z_{so(3)}[^{\gamma}_{\delta}]\over V}Z_2[^0_0]Z_4[^\gamma _\delta]\ , \label{pW4} \end{equation} where $Z_2[^0_0]$ is the contribution of the $T^{(2)}$ compactification, on a $(2,2)$ Lorentzian lattice: $Z_2[^0_0]$=$ \Gamma (2,2)/|\eta|^4$. In the heterotic case, a modular-invariant partition function for $k$ even can be easily obtained using the heterotic map \cite{lls}, \cite{gepner4d}. It consists of replacing in (\ref{pW4}) the $O(4)$ characters associated to the right-moving fermionic coordinates ${\bar\psi}^{\mu}$, with the characters of either $O(12)\otimes E_8$: \begin{equation} (-)^{{\bar\alpha}+{\bar\beta}} \frac{{\bar\vartheta}^{2}[^{\bar\alpha}_{\bar\beta}]}{{\bar\eta}^2} \rightarrow \frac{{\bar\vartheta}^{6}[^{\bar\alpha}_{\bar\beta}]}{{\bar\eta}^6} {1\over 2}\sum_{\gamma\delta=0}^{1}{\bar\vartheta^8[^{\gamma}_{\delta}]\over \bar\eta^8} \label{pfheta} \end{equation} or $O(28)$: \begin{equation} (-)^{{\bar\alpha}+{\bar\beta}} \frac{{\bar\vartheta}^{2}[^{\bar\alpha}_{\bar\beta}]}{{\bar\eta}^2} \rightarrow \frac{{\bar\vartheta}^{14}[^{\bar\alpha}_{\bar\beta}]}{{\bar\eta}^{14}} \ . \label{pfhetb} \end{equation} Other heterotic constructions can be obtain in the case where the extra $\Z_2$ projection acts asymmetrically on the left and right degrees of freedom. In all these constructions the number of space time supersymmetries in flat 4d space ($N=8$ in type II and $N=4$ in heterotic) is reduced by a factor of two when we move in the $W_k$ space. This reduction of space time supersymmetries due to the non trivial mixing of the $SU(2)_k$ characters and those of the internal space can be avoided in the case where the flat construction has a lower number of space time supersymmetries. In order to see how this works, we will examine first the case of $\Z_2$ symmetric orbifold, based on ${\rm I\!R}^4\otimes T^{(2)}\otimes T^{(4)}/\Z_2$, in which the number of supersymmetries is $N=2$ in heterotic and $N=4$ in type II. Contrary to the maximal supersymmetry case, here the number of supersymmetry is already reduced by the $\Z_2$ orbifold projection which acts non trivially to the two spin fields $\tilde{\Theta}_{\pm} = e^{\frac{i}{\sqrt 2}(H^-_1 \pm H^-_2)}\ $ constructed with the $H^-_i$ bosons. The $\Z_2$ orbifold partition function for ${\rm I\!R}^4\otimes T^{(2)}\otimes T^{(4)}/\Z_2$ is: \begin{equation} Z^{\Z_2} = {1\over {\rm Im}\tau|\eta|^4}~{1\over 8} \sum_{\alpha,\beta,{\bar\alpha}\atop {\bar\beta},h,g} (-)^{\alpha+\beta+{\bar\alpha}+{\bar\beta}} \frac{\vartheta^{2}[^{\alpha}_{\beta}]}{\eta^2} \frac{\vartheta^{2}[^{\alpha+h}_{\beta+g}]}{\eta^2}~ \frac{{\bar\vartheta}^{2}[^{\bar\alpha}_{\bar\beta}]}{{\bar\eta}^2} \frac{{\bar\vartheta}^{2} [^{{\bar\alpha}+h}_{{\bar\beta}+g}]}{{\bar\eta}^2} {}~Z_2[^0_0]Z_4[^h _g] \label{pFZ2} \end{equation} The $g$-action projects out in the untwisted sector (h=0) the unwanted spin fields, $\tilde{\Theta}_{\pm}=e^{\frac{i}{\sqrt 2}(H^-_1 \pm H^-_2)}\ $ as usual. Replacing flat space ${\rm I\!R}^4$ with $W_k$ in the orbifold model above, we must specify the extra $\Z_2$ action, which, as we explained in the maximal supersymmetry example, must act non-trivially on the $\tilde{\Theta}_{\pm}=e^{\frac{i}{\sqrt 2}(H^-_1 \pm H^-_2)}\ $ spin fields as well as on the $SU(2)_k$ characters. This action must be in agreement with modular invariance and unitarity. The resulting partition function is: \begin{equation} Z_{W}^{\Z_2} = {{\rm Im}\tau^{1/2}|\eta|^2\over 16} \sum_{\alpha,\beta,{\bar\alpha},{\bar\beta}\atop\gamma,\delta,h,g} (-)^{\alpha+\beta+{\bar\alpha}+{\bar\beta}} \frac{\vartheta^{2}[^{\alpha+\gamma}_{\beta+\delta}]}{\eta^2} \frac{\vartheta^{2}[^{\alpha+h}_{\beta+g}]}{\eta^2} \frac{{\bar\vartheta}^{2}[^{\bar\alpha+\gamma}_{\bar\beta+\delta}]} {{\bar\eta}^2} \frac{{\bar\vartheta}^{2} [^{{\bar\alpha}+h}_{{\bar\beta}+g}]}{{\bar\eta}^2} {Z_{so(3)}[^{\gamma}_{\delta}]\over V} Z_2[^0_0]Z_4[^{h+\gamma} _{g+\delta }] \label{pFZ3} \end{equation} Redefining the parameters $\alpha\rightarrow\alpha-\gamma$, $\beta\rightarrow\beta-\delta$, $h\rightarrow h+\gamma$ $g\rightarrow g+\delta$, the partition function above takes the following factorized form: \begin{equation} Z_{W}^{\Z_2} = {\rm Im}\tau^{1/2}|\eta|^2~{1\over 2}\sum_{\gamma,\delta} {Z_{so(3)}[^{\gamma}_{\delta}]\over V}~{1\over 8} \sum_{\alpha,\beta,{\bar\alpha}\atop {\bar\beta},h,g} (-)^{\alpha+\beta+{\bar\alpha}+{\bar\beta} } \frac{\vartheta^{2}[^{\alpha}_{\beta}]}{\eta^2} \frac{\vartheta^{2}[^{\alpha+h}_{\beta+g}]}{\eta^2} \frac{{\bar\vartheta}^{2}[^{\bar\alpha}_{\bar\beta}]}{{\bar\eta}^2} \frac{{\bar\vartheta}^{2} [^{{\bar\alpha}+h}_{{\bar\beta}+g}]}{{\bar\eta}^2} Z_2[^0_0]Z_4[^{h} _{g}] \label{pFZ2fact} \end{equation} Using the heterotic map (\ref{pfheta}) or (\ref{pfhetb})) we obtain heterotic constructions in curved space with $N=2$ spacetime supersymmetric spectrum. Since the $SO(3)_{k/2}$ contribution factorizes, the number of space-time supersymmetries remains the same as in flat space. The factorization property above, is not a special property of the $\Z_2$ symmetric orbifold but is generic for all ${\rm I\!R}^4\otimes K$ models provided the internal space $K$ is not a theory that produces maximal supersymmetry. Indeed, in cases where the internal $K$ theory is non-trivial, the spin fields which are non-BRST invariant in $W_k\otimes K$ are already absent in the flat space ground state. The validity of this statement can be proven in all orbifold constructions. The reason of the non-factorization in the case where $K$ is toroidal is due to the reduction by a factor of two of the covariantly constant spinors in the $W_k$ background indicating that spacetime supersymmetry cannot be maximal. When the internal space is not trivial then the number of space time supersymmetries is already reduced in flat space and thus, further reduction is not necessary. Observe however, that the odd spin representations of the $SU(2)_k$ ($2j$=odd) are absent, since they are projected out by the $\Z_2$ projection discussed above. Therefore, the correct target space is the $\Z_2$ orbifold of $SU(2)_k$, namely that of $SO(3)_{k/2}$. Thanks to the factorization property described above in the non-maximal supersymmetric case, we can always construct the curved space-time partition function, $Z^W(\tau,{\bar \tau})$ in terms of that of flat space $Z_{0}(\tau,{\bar \tau})$, \begin{equation} Z^W(\tau,{\bar \tau})= {\rm Im}\tau^{3/2}|\eta(\tau)|^6~{\Gamma(SO(3)_{k/2})\over V}Z_{0}(\tau,{\bar \tau})\label{zero} \end{equation} where $\Gamma(SO(3)_{k/2})$ is the partition function of the $SO(3)$ WZW model at level $k/2$: \begin{equation} \Gamma(SO(3)_{k/2})={1\over 2}\sum_{\gamma,\delta=0}^{1}Z_{so(3)}[^{\gamma}_{\delta}] \end{equation} \section{The $\sigma$-model description of magnetic and gravitational backgrounds}\setcounter{equation}{0} The starting 4-d spacetime (we will use Euclidean signature here) is described by the $SO(3)_{k/2}\times {\rm I\!R}_{Q}$ CFT. The heterotic $\sigma$-model that describes this space is\footnote{ In most formulae we set $\alpha'=1$ unless stated otherwise.} \begin{equation} S_{4d}={k\over 4}{\bf I}_{SO(3)}(\alpha,\beta,\gamma) +{1\over 2\pi}\int d^2z \left[\partial x^{0}\bar \partial x^{0}+\psi^{0}\bar \partial\psi^{0}+\sum_{a=1}^{3}\psi^{a}\bar \partial\psi^{a}\right]+ {Q\over 4\pi}\int \sqrt{g}R^{(2)}x^{0} \end{equation} while the SU(2) action can be written in Euler angles as \begin{equation} {\bf I}_{SO(3)}(\alpha,\beta,\gamma)={1\over 2\pi}\int d^2 z\left[\partial\alpha\bar \partial\alpha+\partial\beta\bar \partial\beta +\partial\gamma\bar \partial\gamma+2\cos\beta\partial\alpha\bar \partial\gamma\right] \end{equation} with $\beta\in[0,\pi]$, $\alpha,\gamma\in[0,2\pi]$ and $k$ is a positive even integer. In the type II case we have to add also the right moving fermions $\bar \psi^{i}$, $1\leq i\leq 4$. The fermions are free (this is a property valid for all supersymmetric WZW models). Comparing with the general (bosonic) $\sigma$-model \begin{equation} S={1\over 2\pi}\int d^2 z (G_{\mu\nu}+B_{\mu\nu})\partial x^{\mu}\bar \partial x^{\nu}+ {1\over 4\pi}\int \sqrt{g}R^{(2)}\Phi(x) \end{equation} we can identify the non-zero background fields as \begin{equation} G_{00}=1\;\;,\;\; G_{\alpha\a}=G_{\beta\b}=G_{\gamma\g}={k\over 4} \label{3s1} \end{equation} \begin{equation} G_{\alpha\gamma}={k\over 4}\cos\beta\;\;\;,\;\;\;B_{\alpha\gamma}={k\over 4}\cos\beta \label{3s2}\end{equation} \begin{equation} \Phi=Qx^{0}={x^{0}\over \sqrt{k+2}}\label{dil} \end{equation} where the relation between $Q$ and $k$ is required from the requirement that the (heterotic) central charge should be $(6,4)$, in which case we have (4,0) superconformal invariance, \cite{k}. The perturbation that turns on a chromo-magnetic field in the $\mu=3$ direction is proportional to $(J^{3}+\psi^1\psi^2)\bar J$ where $\bar J$ is a right moving current belonging to the Cartan subalgebra of the heterotic gauge group. It is normalized so that $\langle\bar J(1)\bar J(0)\rangle=k_{g}/2$. Since \begin{equation} J^{3}=k(\partial\gamma+\cos\beta\partial\alpha)\;\;\;,\;\;\;J^{3}=k(\bar \partial\alpha+\cos\beta\bar \partial\gamma) \end{equation} this perturbation changes the $\sigma$-model action in the following way: \begin{equation} \delta S_{4d}={\sqrt{kk_{g}}H\over 2\pi}\int d^2 z(\partial\gamma+\cos\beta\partial\alpha)\bar J \label{gauge} \end{equation} In the type II case $\bar J$ is a bosonic current (it has a left moving partner) and we can easily show that the $\sigma$-model with action $S_{4d}+\delta S_{4d}$ is conformally invariant to all orders in $\alpha'$. This can be seen by writing $\bar J=\bar \partial \phi$ and noticing that \begin{equation} {k\over 4}{\bf I}_{SO(3)}(\alpha,\beta,\gamma)+\delta S_{4d}+{k_{g}\over 4\pi}\int d^2 z~\partial\phi\bar \partial\phi={k\over 4}{\bf I}_{SO(3)}\left(\alpha,\beta,\gamma+2\sqrt{k_{g}\over k}H\phi\right)+ \label{wzw}\end{equation} $$+{k_{g}(1-2H^2)\over 4\pi}\int d^2 z~\partial\phi\bar \partial\phi $$ It is already obvious from (\ref{wzw}) that something special happens at $H^2=1/2$. In fact in the toroidal case that would correspond to a boundary of moduli space. It is the limit Im$U\to \infty$ in the case of a $(2,2)$ lattice. Here the interpretation would be of a maximum magnetic field. We will see more signals of this later on. Reading the spacetime backgrounds from (\ref{gauge}) is not entirely trivial but straightforward. In type II case (which corresponds to standard Kalutza-Klein reduction) the correct metric has an $A_{\mu}A_{\nu}$ term subtracted \cite{dpn}. In the heterotic case there is a similar subtraction but the reason is different. It has to do with the anomaly in the holomorphic factorization of a boson (see for example \cite{p}). The background fields have to be solutions (in leading order in $\alpha '$) to the following equations of motion \cite{eff}: \begin{equation} \delta c={3\over 2}\left[4(\nabla\Phi)^2-{10\over 3}\hbox{{$\sqcup$}\llap{$\sqcap$}}\Phi-{2\over 3}R+{1\over 12g^2}F^{a}_{\mu\nu}F^{a,\mu\nu}\right]=0 \label{1}\end{equation} \begin{equation} R_{\mu\nu}-{1\over 4}H^2_{\mu\nu}-{1\over 2g^2}F^a_{\mu\rho}{F^{a}_{\nu}}^{\rho}+2\nabla_{\mu}\nabla_{\nu} \Phi=0\label{2} \end{equation} \begin{equation} \nabla^{\mu}\left[e^{-2\Phi}H_{\mu\nu\rho}\right]=0\label{3} \end{equation} \begin{equation} \nabla^{\nu}\left[e^{-2\Phi}F^{a}_{\mu\nu}\right] -{1\over 2}F^{a,\nu\rho} H_{\mu\nu\rho}e^{-2\Phi}\label{4} \end{equation} which stem from the variation of the effective action, \begin{equation} S=\int d^{4}x\sqrt{G}e^{-2\Phi}\left[R+4(\nabla\Phi)^2-{1\over 12}H^2-{1\over 4g^2}F^a_{\mu\nu}F^{a,\mu\nu}+{\delta c\over 3}\right]\label{5} \end{equation} where we have displayed a gauge field $A^{a}_{\mu}$, (abelian or non-abelian) and set $g_{\rm string}=1$. The gauge coupling is $g^2=2/k_{g}$ due to the normalization of the currents in (\ref{norm}), \begin{equation} F^a_{\mu\nu}=\partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu}+f^{abc} A^{b}_{\mu}A^{c}_{\nu} \label{6} \end{equation} \begin{equation} H_{\mu\nu\rho}=\partial_{\mu}B_{\nu\rho}-{1\over 2g^2}\left[A^a_{\mu}F^a_{\nu\rho}-{1\over 3}f^{abc}A^{a}_{\mu}A^{b}_{\nu}A^{c}_{\rho}\right]+{\rm cyclic}\;\;{\rm permutations}\label{7} \end{equation} and $f^{abc}$ are the structure constants of the gauge group. In this paper we will restrict ourselves to abelian gauge fields (in the cartan of a non-abelian gauge group). It is not difficult now to read from (\ref{gauge}) the background fields that satisfy the equations above. The non-zero components are: \begin{equation} G_{00}=1\;\;,\;\;G_{\beta\b}={k\over 4}\;\;,\;\;G_{\alpha\gamma}={k\over 4}(1-2H^2)\cos\beta \label{8}\end{equation} \begin{equation} G_{\alpha\a}={k\over 4}(1-2H^2\cos^2\beta)\;\;,\;\;G_{\gamma\g}={k\over 4}(1-2H^2)\;\;,\;\;B_{\alpha\gamma}={k\over 4}\cos\beta \label{9}\end{equation} \begin{equation} A_{a}=g\sqrt{k}H\cos\beta\;\;\;,\;\;\;A_{\gamma}=g\sqrt{k}H \label{10}\end{equation} and the same dilaton as in (\ref{dil}). As shown before this background is exact to all orders in the $\alpha '$ expansion with simple modification $k\to k+2$. It is interesting to note that \begin{equation} \sqrt{{\rm det}G}=\sqrt{1-2H^2}\left({k\over 4}\right)^{3/2}\sin\beta \end{equation} which indicates, as advertised earlier, that something special happens at $H_{max}=1/\sqrt{2}$. At this point the curvature is regular. In fact, this is a boundary point where the states that couple to the magnetic field (i.e. states with non-zero ${\cal{Q}}+I$ and/or $e$) become infinitely massive and decouple. This is the same phenomenon as the degeneration of the Kh\"aler structure on a two-torus (${\rm Im}U\to\infty$). Thus, this point is at the boundary of the magnetic field moduli space. This is very interesting since it implies the existence of a maximal magnetic field \begin{equation} |H|\leq H_{max}={1\over \sqrt{2}} \end{equation} We should note here that the deformation of the spherical geometry by the magnetic field is smooth for all ranges of parameters, even at the boundary point $H=1/\sqrt{2}$. To monitor better the back-reaction of the effective field theory geometry we should first write the three-sphere with the round metric (\ref{3s1}), (\ref{3s2}), as the (Hopf) fibration of $S^1$ as fiber and a two-sphere as base space: \begin{equation} ds^2_{\rm 3-sphere}={k\over 4}\left[ds^2_{\rm 2-sphere}+(d\gamma+\cos\beta d\alpha)^2\right] \label{hopf} \end{equation} with \begin{equation} ds^2_{\rm 2-sphere}=d\beta^2+\sin^2\beta d\alpha^2 \end{equation} The second term in (\ref{hopf}) is the metric of the $S^1$ fiber, and its non-trivial dependence on $\alpha,\beta$ signals the non-triviality of the Hopf fibration. This metric has $SO(3)\times SO(3)$ symmetry. The metric (\ref{8}), (\ref{9}) containing the backreaction to the non-zero magnetic field can be written as \begin{equation} ds^2={k\over 4}\left[ds^2_{\rm 2-sphere}+(1-2H^2)(d\gamma+\cos\beta d\alpha)^2\right] \label{back}\end{equation} It is obvious from (\ref{back}) the magnetic field changes the radius of the fiber and breaks the $SO(3)\times SO(3)$ symmetry to the diagonal $SO(3)$. It is also obvious that at $H=1/\sqrt{2}$, the radius of the fiber becomes zero. All the curvature invariants are smooth (and constant due to the $SO(3)$ symmetry) As mentioned in the introduction, we have another marginal deformation (\ref{gravi}) which turns on curvature as well as antisymmetric tensor and dilaton. The essential part of this perturbation is the $J^{3}\bar J^3$ part which deforms the Cartan torus of SO(3) and the exact bosonic $\sigma$-model action was given in \cite{hs,gk}. We will use this result to derive the background fields associated with both gauge and gravitational deformation. After some algebra we obtain \def\lambda{\lambda} \begin{equation} G_{00}=1\;\;\;G_{\beta\b}={k\over 4}\label{11} \end{equation} \begin{equation} G_{\alpha\a}={k\over 4}{(\lambda^2+1)^2-(8H^2\lambda^2+(\lambda^2-1)^2)\cos^2\beta\over (\lambda^2+1+(\lambda^2-1)\cos\beta)^2} \label{14}\end{equation} \begin{equation} G_{\gamma\g}={k\over 4}{(\lambda^2+1)^2-8H^2\lambda^2-(\lambda^2-1)^2\cos^2\beta\over (\lambda^2+1+(\lambda^2-1)\cos\beta)^2}\label{15}\end{equation} \begin{equation} G_{\alpha\gamma}={k\over 4}{4\lambda^2(1-2H^2)\cos\beta+(\lambda^4-1)\sin^2\beta\over (\lambda^2+1+(\lambda^2-1)\cos\beta)^2}\label{16} \end{equation} \begin{equation} B_{\alpha\gamma}={k\over 4}{\lambda^2-1+(\lambda^2+1)\cos\beta\over (\lambda^2+1+(\lambda^2-1)\cos\beta)} \label{17} \end{equation} \begin{equation} A_{a}=2g\sqrt{k}{H\lambda\cos\beta\over (\lambda^2+1+(\lambda^2-1)\cos\beta)}\label{18} \end{equation} \begin{equation} A_{\gamma}=2g\sqrt{k}{H\lambda\over (\lambda^2+1+(\lambda^2-1)\cos\beta)}\label{19}\end{equation} \begin{equation} \Phi={t\over \sqrt{k+2}}-{1\over 2}\log\left[\lambda+{1\over \lambda}+\left(\lambda-{1\over \lambda} \right)\cos\beta\right]\label{20} \end{equation} It is straightforward to verify that the fields above solve the equations of the effective field theory. We now have an additional modulus which governs the gravitational perturbation, namely $\lambda$ which we can take it to be a non-negative real number. There are however duality symmetries that act on the moduli $H,\lambda$. The first is a $Z_{2}^{I}$ duality symmetry $\lambda\to 1/\lambda$ \cite{gk} accompanied by a reparamerization $\beta\to \pi-\beta$, $\cos\beta\to -\cos\beta$ and $\alpha\to -\alpha$ under which the background fields and thus the CFT are invariant. This is a parity-like symmetry (in the $\alpha$ direction) since if we do not transform $\alpha$ then $A_{a}\to -A_{a}$ and $A_{\gamma}\to A_{g}$. There is another $Z_{2}^{II}$ duality which acts on $H$ as $H\to -H$. This a charge conjugation symmetry since it changes the sign of the gauge fields. The combined transformation is a CP symmetry since it changes the sign of the Lorentz generator (in the third direction) $J^{3}+\bar J^{3}$. Again here the deformed geometry is smooth for all values of $H,\lambda$ except at the boundaries $\lambda=0,\infty$ where the magnetic field turns off and the three-sphere degenerates to the (classically) singular geometry of the $R\times SU(2)_{k}/U(1)$, \cite{gk}. \section{Conformal Field Theory description of magnetic and gravitational backgrounds} \setcounter{equation}{0} Our aim in this section is to define the deformation of the original string ground state, that turns on magnetic fields and curvature, and study the exact spectrum. In particular we find the presence of instabilities of the tachyonic type associated to such backgrounds. We will focus on heterotic 4-D string ground states, described in detail in the previous section, although the extension to type II ground states is straightforward. \def{\cal Q}{{\cal Q}} As mentioned in the introduction, the vertex operator which turns on a chromo-magnetic field background $B^{a}_{i}$ is \begin{equation} V^{a}_{i}=(J^i+{1\over 2}\epsilon^{i,j,k}\psi^j\psi^k)\bar J^{a} \end{equation} This vertex operators is of the current-current type. In order for such perturbations to be marginal (equivalently the background to satisfy the string equations of motion) we need to pick a single index $i$, which we choose to be $i=3$ and need to restrict the gauge group index $a$ to be in the Cartan of the gauge group. We will normalize the antiholomorphic currents $\bar J^a$ in each simple or U(1) component $G_{i}$ of the gauge group $G$ as \begin{equation} \langle \bar J^a(\bar z)\bar J^b(0)\rangle ={k_{i}\over 2}{\delta^{ab}\over \bar z^2}\label{norm} \end{equation} With this normalization, the field theory gauge coupling is $g_{i}^2=2/k_{i}$. Thus the most general (marginal) chromo-magnetic field is generated from the following vertex operator \begin{equation} V_{magn}={(J^3+\psi^1\psi^2)\over \sqrt{k+2}}{{\vec F_{i}}\cdot{\bar {\vec J_{i}}}\over \sqrt{k_{i}}}\label{magn} \end{equation} where the index $i$ labels the simple or $U(1)$ components $G_{i}$ of the gauge group and $\bar{\vec J_{i}}$ is a $r_{i}$-dimensional vector of currents in the Cartan of the group $G_{i}$ ($r_{i}$ is the rank of $G_{i}$). The repeated index $i$ implies summation over the simple components of the gauge group. We would like to obtain the exact one-loop partition function in the presence of such perturbation. Since this is an abelian current-current perturbation, the deformed partition function can be obtained by an $O(1,\sum_{i}r_{i})$ boost of the charged lattice of the undeformed partition function, computed in the previous section. We will indicate the method in the case where we turn on a single magnetic field $F$, a gauge group factor with central element $k_{g}$, in which case \begin{equation} V_{F}= F{(J^3+\psi^1\psi^2)\over \sqrt{k+2}}{\bar J\over \sqrt{k_{g}}} \end{equation} Let us denote by ${\cal Q}$ the zero mode of the holomorphic helicity current $\psi^1\psi^2$, $\bar {\cal P}$ the zero mode of the antiholomorphic current $\bar J$ and $I,\bar I$ the zero modes of the $SU(2)$ currents $J^{3},\bar J^{3}$ respectively. Then, the relevant parts of $L_{0}$ and $\bar L_{0}$ are \begin{equation} L_{0}={{\cal Q}^2\over 2}+{I^2\over k}+\cdots\;\;\;,\;\;\;\bar L_{0}={\bar {\cal P}^2\over k_{g}}+\cdots\label{k1} \end{equation} We will rewrite $L_{0}$ as \begin{equation} L_{0}={({\cal Q}+I)^2\over k+2}+{k\over 2(k+2)}\left({\cal Q}-{2\over k}I\right)^2+\cdots \end{equation} where we have separated the relevant supersymmetric zero mode ${\cal Q}+I$ and its orthogonal complement ${\cal Q}-2I/k$ which will be a neutral spectator to the perturbing process. What remains to be done is an $O(1,1)$ boost that mixes the holomorphic current ${\cal Q}+I$ and the antiholomorphic one $\bar {\cal P}$. This is straightforward with the result \begin{equation} L_{0}'= {k\over 2(k+2)}\left({\cal Q}-{2\over k}I\right)^2+\left(\cosh x {{\cal Q}+I\over \sqrt{k+2}}+\sinh x {\bar {\cal P}\over \sqrt{k_{g}}}\right)^2+\cdots \label{k2} \end{equation} \begin{equation} \bar L_{0}'= \left(\sinh x {{\cal Q}+I\over \sqrt{k+2}}+\cosh x {\bar {\cal P}\over \sqrt{k_{g}}}\right)^2+\cdots\label{k3} \end{equation} where $x$ is the parameter of the $O(1,1)$ boost. Thus we obtain from (\ref{k2}), (\ref{k3}) the change of $L_{0}$, $\bar L_{0}$ as \begin{equation} \delta L_{0}\equiv L_{0}'-L_{0}=\delta \bar L_{0}\equiv \bar L_{0}'-\bar L_{0}= F{({\cal Q}+I)\over \sqrt{k+2}}{\bar {\cal P}\over \sqrt{k_{g}}}+{\sqrt{1+F^2}-1\over 2}\left[ {({\cal Q}+I)^2\over k+2}+{\bar {\cal P}^2\over k_{g}}\right] \end{equation} where we have identified \begin{equation} F\equiv \sinh (2x) \end{equation} we are now able to compute with the more general marginal perturbation which is a sum of the general magnetic perturbation (\ref{magn}) and the gravitational perturbation \begin{equation} V_{grav}= {\cal R} {(J^3+\psi^1\psi^2)\over \sqrt{k+2}}{\bar J^3\over \sqrt{k}} \end{equation} The only extra ingredient we need is an $O(1+\sum_{i}r_{i})$ transformation to mix the antiholomorphic currents. Thus, we obtain \begin{equation} \delta L_{0}=\delta \bar L_{0}=\left[{{\cal R}\bar I\over \sqrt{k}}+{{\vec F}_{i}\cdot {\bar {\vec {\cal P}}}_{i}\over \sqrt{k_{i}}}\right]{({\cal Q}+I)\over \sqrt{k+2}}+\label{gener} \end{equation} $$+{\sqrt{1+{\cal R}^2+{\vec F}_{i}\cdot {\vec F}_{i}}-1\over 2}\left[ {({\cal Q}+I)^2\over k+2}+({\cal R}^2+{\vec F}_{i}\cdot {\vec F}_{i})^{-1}\left( {{\cal R}\bar I\over \sqrt{k}}+{{\vec F}_{i}\cdot {\bar {\vec {\cal P}}}_{i}\over \sqrt{k_{i}}}\right)^2\right] $$ {}From now on we focus in the case where we have a single chromo-magnetic field $F$ as well as the curvature perturbation ${\cal R}$. Then (\ref{gener}) simplifies to \begin{equation} \delta L_{0}=\delta \bar L_{0}= \left[{\cal R}{\bar I\over \sqrt{k}}+F{\bar {\cal P}\over \sqrt{k_{g}}}\right]{({\cal Q}+I)\over \sqrt{k+2}}+\label{gener2} \end{equation} $$+{\sqrt{1+{\cal R}^2+F^2}-1\over 2}\left[ {({\cal Q}+I)^2\over k+2}+({\cal R}^2+F^2)^{-1}\left( {\cal R}{\bar I\over \sqrt{k}}+F{\bar {\cal P}\over \sqrt{k_{g}}}\right)^2\right] $$ Eq. (\ref{gener2}) can be written in the following form which will be useful in order to compare with the field theory limit \begin{equation} \delta L_{0}={1+\sqrt{1+F^2+{\cal R}^2}\over 2}\left[{({\cal Q}+I)\over \sqrt{k+2}}+{1\over 1+\sqrt{1+F^2+{\cal R}^2}}\left({\cal R}{\bar I\over \sqrt{k}} +F{\bar {\cal P}\over \sqrt{k_{g}}}\right)\right]^2 \end{equation} $$-{({\cal Q}+I)^2\over k+2} $$ and for ${\cal R}=0$ as \begin{equation} \delta L_{0}={1+\sqrt{1+F^2}\over 2}\left[{({\cal Q}+I)\over \sqrt{k+2}}+{F\over 1+\sqrt{1+F^2}}{\bar {\cal P}\over \sqrt{k_{g}}}\right]^2 -{({\cal Q}+I)^2\over k+2}\label{k10} \end{equation} Eq. (\ref{zero}) along with (\ref{gener}) provide the complete and exact spectrum of string theory in the presence of the chromo-magnetic fields $\vec F_{i}$ and curvature ${\cal R}$. We will analyse first the case of a single magnetic field $F$ and use (\ref{k10}). Let $L_{0}=M_{L}^2$ and $\bar L_{0}=M_{R}^2$. Then \begin{equation} M^2_{L}=-{1\over 2}+{{\cal Q}^2\over 2}+{1\over 2}\sum_{i=1}^{3}{\cal Q}_{i}^2+{(j+1/2)^2-({\cal Q}+I)^2\over k+2}+E_{0}+\label{k11} \end{equation} $$+{1+\sqrt{1+F^2}\over 2}\left[{({\cal Q}+I)\over \sqrt{k+2}}+{F\over 1+\sqrt{1+F^2}}{\bar {\cal P}\over \sqrt{k_{g}}}\right]^2 $$ \begin{equation} M^2_{R}=-1+{\bar {\cal P}^2\over k_{g}}+{(j+1/2)^2-({\cal Q}+I)^2\over k+2}+\bar E_{0}+ \end{equation} $$+{1+\sqrt{1+F^2}\over 2}\left[{({\cal Q}+I)\over \sqrt{k+2}}+{F\over 1+\sqrt{1+F^2}}{\bar {\cal P}\over \sqrt{k_{g}}}\right]^2 $$ where, the $-1/2$ is the universal intercept in the N=1 side, ${\cal Q}_{i}$ are the internal helicity operators (associated to the internal left-moving fermions), $E_{0},\bar E_{0}$ contain the oscillator contributions as well as the internal lattice (or twisted) contributions, and $j=0,1,2,\cdots,k/2$\footnote{Remember that $k$ is an even integer for $SO(3)$.}, $j\geq |I|\in Z$. We can see here another reason for the need of the SO(3) projection. We do not want half integral values of $I$ to change the half-integrality of the spacetime helicity ${\cal Q}$. Since for physical states $L_{0}=\bar L_{0}$ it is enough to look at $M_{L}^2$ which in our conventions is the side that has $N=1$ superconformal symmetry. Let us consider first at how the low lying spectrum of space-time fermions is modified. For this we have to take ${\cal Q}={\cal Q}_{i}=\pm 1/2$. Then $M_{L}^2$ can be written as a sum of positive factors, $E_{0}\geq 0$, $(j+1/2)^2\geq (\pm 1/2+I)^2$ and \begin{equation} {1+\sqrt{1+F^2}\over 2}\left[{({\cal Q}+I)\over \sqrt{k+2}}+{F\over 1+\sqrt{1+F^2}}{\bar {\cal P}\over \sqrt{k_{g}}}\right]^2 \geq 0\label{k14} \end{equation} Thus fermions cannot become tachyonic and this a good consistency check for our spectrum since a ``tachyonic" fermion is a ghost. This argument can be generalized to all spacetime fermions in the theory. Bosonic states can become tachyonic though, but for this to happen, as in field theory they need to have non-zero helicity. Since $(j+1/2)^2\geq I^2$ and $E_{0}\geq 0$, a state needs a non-zero value for ${\cal Q}$ and the minimum possible value for ${\cal Q}_{i}^2$ (consistent with the GSO projection) as well as $E_{0}=0$ in order to have a chance to become tachyonic. Also we need $j=\pm I$ and ${\cal Q} I$ positive. For such states, imposing $L_{0}=\bar L_{0}$ we obtain \begin{equation} {\cal Q}^2-{2\over k_{g}}\bar {\cal P}^2+1=2\bar E_{0}\geq 0\label{k16} \end{equation} and thus the minimal value for $M_{L}^2$ can be written as \begin{equation} M^2_{min}={{\cal Q}^2-1\over 2}+{(|I|+1/2)^2-({\cal Q}+I)^2\over k+2}+ {1+\sqrt{1+F^2}\over 2}\left[{({\cal Q}+I)\over \sqrt{k+2}}+{F\over 1+\sqrt{1+F^2}}{\bar {\cal P}\over \sqrt{k_{g}}}\right]^2 \label{k13} \end{equation} and due to Eq. (\ref{k14}) and the fact that $2I\leq k$ we obtain \begin{equation} k|{\cal Q}|(|{\cal Q}|-2)\leq 3/2\;\;\;,\;\;\;|{\cal Q}|=1,2,\cdots\label{k24} \end{equation} \begin{equation} 2|I|\geq {-k{\cal Q}^2+k+3/2\over 1-2|{\cal Q}|}\label{k25} \end{equation} which imply that either $|{\cal Q}|=1$ and $|I|=0,1,\cdots,k/2$, or $|{\cal Q}|=2$ and $|I|=k/2$, provided $k>0$. However, due to the GSO projection, ${\cal Q}$ mast be an odd integer. Thus, for $k>0$ instabilities are due to helicity $\pm 1$ particles. Let us introduce the variables \begin{equation} H={F\over \sqrt{2}(1+\sqrt{1+F^2})}\;\;\;,\;\;\;e=\sqrt{2\over k_{g}}\bar {\cal P} \end{equation} $H$ is the natural magnetic field from the $\sigma$-model point of view (see section 3) and $e$ is the charge. Notice that while $F$ varies along the whole real line, $|H|\leq 1/\sqrt{2}$. At $H_{max}=1/\sqrt{2}$ we can see from (\ref{k11}) that there is an infinite number of states whose mass becomes zero, so it is a decompactification boundary. Eq. (\ref{k16}) can be rewritten as \begin{equation} e^2\leq {\cal Q}^2+1\label{k17} \end{equation} Then, there are tachyons provided \begin{equation} {1\over 1-2H^2}\left({({\cal Q}+I)\over \sqrt{k+2}}+eH\right)^2+{{\cal Q}^2-1\over 2}+{(|I|+1/2)^2-({\cal Q}+I)^2\over k+2}\leq 0\label{tach} \end{equation} For (\ref{tach}) to have solutions we must have \begin{equation} e^2\geq {\cal Q}^2-1+2{(|I|+1/2)^2\over k+2}\label{k18} \end{equation} which along with (\ref{k17}) implies that $(|I|+1/2)^2\leq k+2$. It is not difficult to see that the first instability sets in, induced from the $I=0$ states. There is also an upper critical magnetic field beyond which no state is tachyonic. This is obtained by considering the largest possible value for $|I|$ (compatible with $(|I|+1/2)^2\leq k+2$). We will leave the charge free for the moment although there are certainly constraints on it depending on the gauge group. For example for the $E_{6}$ or $E_{8}$ groups we have $e^2_{min}=1/4$, and for all realistic non-abelian gauge groups $e_{min}={\cal O}(1)$. For toroidal $U(1)$'s however $e_{min}$ can become arbitrarily small by tuning the parameters of the torus. Note however that in any case for the potential tachyonic states with $|{\cal Q}|=1$ the charge must satisfy \begin{equation} {1\over 2(k+2)} \leq e^2 \leq 2\label{k20} \end{equation} Thus for $|{\cal Q}|=1$ we obtain the presence of tachyons provided that \begin{equation} H^{\rm crit}_{\rm min}\leq |H| \leq H^{\rm crit}_{\rm max} \end{equation} with \begin{equation} H^{\rm crit}_{\rm min}={\mu\over |e|} {1-{\sqrt{3}\over 2}\sqrt{1-{1\over 2}\left({\mu\over e}\right)^2}\over 1+{3\over 2}\left({\mu\over e}\right)^2} \end{equation} \begin{equation} H^{\rm crit}_{\rm max}={\mu\over |e|}{J+1+ \sqrt{\left(J+{3\over 4}\right)\left(1-2\left(J+{1\over 2}\right)^2{\mu^2\over e^2}\right)}\over 1+\left(2J+{3\over 2}\right){\mu^2\over e^2}} \label{q1} \end{equation} where \begin{equation} J={\rm integral ~~part~~of~~~}-{1\over 2}+{|e|\over \sqrt{2}\mu} \end{equation} We have also introduced the IR cutoff scale $k+2=1/\mu^2$. We note that for small $\mu$ and $|e|\sim {\cal O}(1)$ $H^{\rm crit}_{\rm min}$ is of order ${\cal O}(\mu)$. However $H^{\rm crit}_{\rm max}$ is below $H_{\rm max}=1/\sqrt{2}$ by an amount of order ${\cal O}(\mu)$. Thus for small values of $H$ there are no tachyons until a critical value $H^{\rm crit}_{\rm min}$ where the theory becomes unstable. For $|H|\geq H^{\rm crit}_{\rm max}$ the theory is stable again till the boundary $H=1/\sqrt{2}$. It is interesting to note that if there is a charge in the theory with the value $|e|=\sqrt{2}\mu$ then $H^{\rm crit}_{\rm max}=1/\sqrt{2}$ so there is no region of stability for large magnetic fields. For small $\mu$ there are always charges satisfying (\ref{k20}) which implies that there is always a magnetic instability. However even for $\mu={\cal O}(1)$ the magnetic instability is present for standard gauge groups that have been considered in string model building (provided it has charged states in the perturbative spectrum). The behavior above should be compared to the field theory behavior (\ref{class}). There we have an instability provided there is a particle with $gS\geq 1$. Then the theory is unstable for \begin{equation} |H|\geq {M^2\over |e|(gS-1)} \end{equation} where $M$ is the mass of the particle (or the mass gap). However there is no restauration of stability for large values of $H$. This happens in string theory due to the backreaction of gravity. There is also another difference. In field theory $H_{crit}\sim \mu^2$ while in string theory $H_{crit}\sim \mu M_{\rm planck}$ where we denoted by $\mu$ the mass gap in both cases and $M^2_{str}=1/\alpha'g^2_{\rm string}$. We will also study the special case $k=0$, which was left out from the analysis above. This corresponds to a strongly curved spacetime (the curvature of $S^3$ is of the order of the string scale). We know of course from the CFT that for $k=0$ the $S^3$ decouples (only the ground state is left). This is a non-critical string theory since from the bosonic part of 4-d, only the Liouville field survives. Moreover, $H$ loses its meaning as a magnetic field (since it couples only to the helicity operator). In this case all (odd) helicity states can become tachyonic and we obtain an instability for \begin{equation} H_{min}^{k=0}\leq H^{k=0} \leq H_{max}^{k=0} \end{equation} with \begin{equation} H^{k=0}_{\rm max}=H_{\rm max}={1\over \sqrt{2}} \end{equation} \begin{equation} H^{k=0}_{\rm min}={1\over \sqrt{2}}{|e|-\sqrt{3(e^2-1/4)/4}\over e^2+3/4} \end{equation} The first tachyonic instability related to $H^{k=0}_{\rm min}$ is induced by $|{\cal Q}|=1$ states. The theory never becomes stable again since for all $H\leq H_{\rm max}$ there are tachyonic states for arbitrary high values of $|{\cal Q}|$. The analysis above applies to magnetic fields embedded in non-abelian gauge groups, not broken by the conventional Higgs effect. We will also consider however broken non-abelian gauge groups. Consider the internal CFT containing a circle of radius R. For arbitrary values of $R\not= 1$ there is a U(1) gauge symmetry which is enhanced at $R=1$ to $SU(2)$. The $W^{\pm}$ bosons have masses proportional to $(R-1/R)^2$ and become massless at $R=1$. In such a case we will again consider states with ${\cal{Q}}=1$, ${\cal{Q}}_{i}=0$, $E_{0}=(R-1/R)^2/4$ and $\bar {\cal P}/\sqrt{k_{g}}=(R+1/R)/2$ in (\ref{k11}). The condition for the W-bosons becoming tachyonic is \begin{equation} (1-2H^2)\left[{1\over 4}\left(R-{1\over R}\right)^2-\left(|I|+{3\over 4}\right) \mu^2\right]+\left[(|I|+1)\mu+{H\over \sqrt{2}}\left(R+{1\over R}\right)\right]^2\leq 0 \label{u1}\end{equation} It is obvious that the first factor has to be negative so that there tachyons provided \begin{equation} {1\over 4\mu^2}\left(R-{1\over R}\right)^2-{3\over 4}\leq |I|\leq {1\over 2\mu^2}-1 \end{equation} from which we obtain \begin{equation} {4-\mu^2-\sqrt{(4-\mu^2)^2-4}\over 2}\leq R^2 \leq {4-\mu^2+\sqrt{(4-\mu^2)^2-4}\over 2} \end{equation} Note that this condition is duality invariant. Again here we have two critical values for the magnetic field as before that can be computed from (\ref{u1}). However there is no instability in the flat limit $\mu\to 0$ unlike the field theory case (\ref{sta}) due to the gravitational back reaction. Let us now study the gravitational perturbation. Using (\ref{gener2}) the mass formula is (in analogy with (\ref{k11}) \begin{equation} M^2_{L}=-{1\over 2}+{{\cal Q}^2\over 2}+{1\over 2}\sum_{i=1}^{3}{\cal Q}_{i}^2+{(j+1/2)^2-({\cal Q}+I)^2\over k+2}+E_{0}+\label{k22} \end{equation} $$+{1+\sqrt{1+{\cal R}^2}\over 2}\left[{({\cal Q}+I)\over \sqrt{k+2}}+{{\cal R}\over 1+\sqrt{1+{\cal R}^2}}{\bar I\over \sqrt{k}}\right]^2 $$ Introducing the $\sigma$-model variable \begin{equation} \lambda=\sqrt{{\cal R} +\sqrt{1+{\cal R}^2}}\;\;\;,\;\;\;{1\over \lambda}=\sqrt{-{\cal R} +\sqrt{1+{\cal R}^2}} \end{equation} (\ref{k22}) becomes \begin{equation} M^2_{L}=-{1\over 2}+{{\cal Q}^2\over 2}+{1\over 2}\sum_{i=1}^{3}{\cal Q}_{i}^2+{(j+1/2)^2-({\cal Q}+I)^2\over k+2}+E_{0}+\label{k23} \end{equation} $$+{1\over 4}\left[\left(\lambda+{1\over \lambda}\right){({\cal Q}+I)\over \sqrt{k+2}}+\left(\lambda-{1\over \lambda}\right){\bar I\over \sqrt{k}}\right]^2 $$ Eqs (\ref{k24}) and (\ref{k25}) are still applicable here, which means that we have to examine only $|{\cal Q}|=1$ and $|I|=0,1,\cdots,k/2$, or $|{\cal Q}|=2$ and $|I|=k/2$. Again $j=|I|$ and $I{\cal Q} >0$. Due to the $\lambda\to 1/\lambda$ duality we will restrict ourselves to the region $\lambda\leq 1$. Thus, the condition for existence of tachyons is \begin{equation} {1\over 4}\left[\left(\lambda+{1\over \lambda}\right){({\cal Q}+I)\over \sqrt{k+2}}+\left(\lambda-{1\over \lambda}\right){\bar I\over \sqrt{k}}\right]^2+ {{\cal Q}^2-1\over 2}+{(|I|+1/2)^2-({\cal Q}+I)^2\over k+2}\leq 0\label{tach2} \end{equation} In order that (\ref{tach2}) have solutions we must have \begin{equation} \left[{{\cal Q}^2-1\over 2}+{(|I|+1/2)^2-({\cal Q}+I)^2\over k+2}\right] \left[{{\cal Q}^2-1\over 2}+{(|I|+1/2)^2-\bar I^2\over k}\right]\geq 0 \label{k26} \end{equation} The first factor was arranged already to be negative so we must ensure that the second factor is also negative. This is impossible for $|{\cal Q}|=2$. Thus we are left with $|{\cal Q}|=1$ and \begin{equation} |\bar I|\geq \sqrt{k\over k+2}\left(|I|+{1\over 2}\right) \label{k27} \end{equation} Thus the state with quantum numbers $(I,\bar I)$ satisfying (\ref{k27}) becomes tachyonic when \begin{equation} \lambda^2_{\rm min}\leq \lambda^2 \leq \lambda^2_{\rm max} \end{equation} with \begin{equation} \lambda^2_{\rm max}={{\bar I^2\over k}-{I^2-1/2\over k+2}+\sqrt{{(I+3/4)\over k+2}\left( {\bar I^2\over k}-{(I+1/2)^2\over k+2}\right)}\over \left({I\over \sqrt{k+2}} +{\bar I\over \sqrt{k}}\right)^2} \end{equation} \begin{equation} \lambda^2_{\rm min}={{\bar I^2\over k}-{I^2-1/2\over k+2}-\sqrt{{(I+3/4)\over k+2}\left( {\bar I^2\over k}-{(I+1/2)^2\over k+2}\right)}\over \left({I\over \sqrt{k+2}} +{\bar I\over \sqrt{k}}\right)^2} \end{equation} For large $k$, $\lambda_{\rm max}$ approaches one, however at the same time the instability region shrinks to zero so that in the limit $\lambda=1,k=\infty$ flat space is stable. \section{The Flat Space Limit}\setcounter{equation}{0} As mentioned earlier, in the limit $k\to \infty$ the 4-d space becomes flat (with zero dilaton). We would like to understand the nature of the magnetic fields in this limit. As a warm-up we will describe first (in the context of field theory) the case of a constant magnetic field in flat space as a limit of a monopole field of a two-sphere in the limit that the radius of the sphere becomes large. Let $g$ be the strength of the monopole field. Then \begin{equation} \vec B_{monopole}={g}{{\vec r}\over r^3}\label{33} \end{equation} We have the Dirac quantization condition for $g$ in terms of the elementary electric charge $e$: $eg=N$ where $N$ is an arbitrary positive integer or half-integer. Let us now consider a charged spinless particle of charge $e$ constrained to move on a two-sphere around the origin, of radius $R$. The (non-relativistic\footnote{The relativistic case is similar up to the zero point shift $m_{0}$ of the energy, a scaling of the Landau spectrum by $1/m_{0}$ and ${\cal O}(m_{0}^{-3})$ relativistic corrections.}) spectrum of such a particle is known \cite{col}: \begin{equation} \Delta E_{j}={1\over R^2}\left[j(j+1)-N^2\right] \label{34}\end{equation} where $j=N,N+1,\cdots$ and $N=eg\in Z/2$. For each $j$ there are $2j+1$ degenerate states. If we define $n=j-N$ then \begin{equation} \Delta E_{n}={1\over R^2}\left[n(n+1)+N(2n+1)\right]\;\;\;,\;\;\;n=0,1,\cdots\label{35} \end{equation} We would like now to take $R\to \infty$. There are two possible limits to consider. The first is the limit where the magnetic flux per unit area is finite. Since the area of the sphere becomes infinite in the flat limit we will have to take the monopole strength to $\infty$ as $g=HR^2+{\cal O}(R)$ where $H$ is the flat space magnetic field. Then $eg=N=eHR^2+{\cal O}(R)$ and \begin{equation} \Delta E^{2d-flat\;\;space}_{n}=eH(2n+1)+{\cal O}(R^{-1})\;\;\;,\;\;\;n=0,1,\cdots \label{36} \end{equation} We have thus recovered the usual formula where $n$ labels the Landau levels and we should remember that each Landau level is infinitely degenerate corresponding to different values of the angular momentum in the direction perpendicular to the plane. The other limit is to keep the monopole strength fixed. In this case we end up with a zero flat space magnetic field and continuous spectrum $E=p^2$ corresponding, not surprisingly, to a free particle in 2-d. Let us consider now (again in the context of field theory) a magnetic field on a three-sphere of radius $R$. In Euler angles: \begin{equation} A_{\alpha}=H\cos\beta\;\;,\;\;A_{\beta}=0\;\;,\;\;A_{\gamma}=H \end{equation} which is exactly the same as the stringy background (\ref{10}) we have found earlier. We will find again the energy spectrum of a particle of electric charge $e$ moving on $S^3$. The Hamiltonian is as usual \begin{equation} \hat {\bf H}={1\over \sqrt{{\rm det} G}}\left(\partial_{\mu}-ieA_{\mu}\right) \sqrt{{\rm det}G}G^{\mu\nu}\left(\partial_{\nu}-ieA_{\nu}\right) \label{38} \end{equation} where $G_{\mu\nu}$ is the round metric on $S^{3}$. Notice that it is different from the stringy metric (\ref{8}),(\ref{9}) which contains the gravitational backreaction. It is straightforward to work out the spectrum of $\hat {\bf H}$ with the result: \begin{equation} \Delta E_{j,m}={1\over R^2}\left[j(j+1)-m^2+(eH-m)^2\right]\label{39} \end{equation} where for $SO(3)$ $j\in Z$ and $-j\leq m\leq j$. We can always parametrize $j,m$ as $j=|m|+n$ with $|m|=0,1,\cdots$ and $n=0,1,2,\cdots$ so \begin{equation} \Delta E_{n,m}={1\over R^2}\left[n(n+1)+|m|(2n+1)\right]+\left({eH-m\over R}\right)^2 \label{40}\end{equation} In order to take the flat space limit and recover Landau levels we have to scale $eH=e\tilde H R^2+\kappa R+{\cal O}(1)$ and $m=e\tilde H R^2+(p_{3}+\kappa) R+{\cal O}(1)$. Then we obtain from (\ref{40}) \begin{equation} \Delta E_{n,p_{3}}=e\tilde H(2n+1)+p_{3}^2+{\cal O}(R^{-1}) \label{41}\end{equation} which is the standard Landau spectrum in 3-d flat space. This reproduces (\ref{class}) for spinless particles, $S=0$. In the discussion above we did not include the gravitational backreaction since we had a round metric for $S^{3}$. This is what we are going to do now. We will start from the background (\ref{8}),(\ref{9}),(\ref{10}) and compute the energy eigenvalues of the (field theory) Hamiltonian given by (\ref{38}). This is straightforward with the answer: \begin{equation} \Delta E_{j,m}={1\over R^2}\left[j(j+1)-m^2+{(eHR-m)^2\over (1-2H^2)}\right]\label{42} \end{equation} and after parametrizing again $j=|m|+n$ with $|m|=0,1/2,1,\cdots$ and $n=0,1,2,\cdots$ we obtain \begin{equation} \Delta E_{n,m}={1\over R^2}\left[n(n+1)+|m|(2n+1)\right]+\left({eHR-m\over R\sqrt{1-2H^2}}\right)^2 \label{43}\end{equation} Notice that the only difference from (\ref{40}) (apart from the different scaling of H which is a convention) is the extra $1-2H^2$ factor in the denominator of the last term. This factor however makes the flat limit quite different. In fact we can see that in order to have Landau levels we have to take $m\sim {\cal O}(R^2)$, in which case we are forced to have from the last term that $H\sim {\cal O}(R)$ in which case the denominator gives a negative contribution. This is obvious from the fact that since there is a maximal value for H we cannot take it to scale as the radius R. So Landau levels disappear from the low energy spectrum, and with $m=pR+{\cal O}(1)$ we obtain in the limit $R\to\infty$ \begin{equation} \Delta E_{p}={(p-eH)^2\over 1-2H^2}+{\cal O}(R^{-1}) \label{44} \end{equation} This implies that the flat space limit is quite different from standard field theory. We have already seen in the previous section that $W$ bosons do not become tachyonic in the flat limit. The field theory spectrum parallels the exact string spectrum in the presence of a magnetic field. The correct identification there is \begin{equation} R\to k+2\;\;\;,\;\;\;m\to Q+J^3\;\;\;,\;\;\; e\to \sqrt{2\over k_{g}}\bar Q \label{450} \end{equation} \begin{equation} H\to {F\over \sqrt{2}(1+\sqrt{1+F^2})}={1\over 2\sqrt{2}}\left[F-{F^3\over 4}+ {\cal O}(F^5)\right] \end{equation} In terms of the CFT variable $F$ the maximum magnetic field $H_{max}=1/\sqrt{2}$ corresponds to the limit $F\to \pm\infty$. As we have seen already the tachyonic instabilities appear before the magnetic field reaches its maximum value. We can now discuss the spectra of particles with spin. For particles that inherit their spin from the helicity operator (this includes massless fermions, and heterotic gauge fields) we can set $S=Q$ and using the string identification (\ref{450}) we obtain the following spectrum \begin{equation} \Delta E_{j,m,S}={1\over k+2}\left[j(j+1)-(m+S)^2+{(eHR-m-S)^2\over (1-2H^2)}\right]\label{422} \end{equation} which to linear order in the magnetic field becomes \begin{equation} \Delta E_{j,m,S}={j(j+1)\over k+2}-{2eH\over \sqrt{k+2}}(m+S)+{\cal O}(H^2) \end{equation} which indicates the possibility of tachyons. The existence of tachyonic modes with non-zero spin was verified in section 4. In a similar manner we can compute the (scalar) field theory spectrum in the combined magnetic and gravitational background (\ref{11})-(\ref{20}) \begin{equation} \Delta E_{j,m,\bar m}= {1\over R^2}\left[j(j+1)-m^2+{(2eHR-(\lambda+1/\lambda)m-(\lambda-1/\lambda)\bar m)^2\over 4(1-2H^2)}\right] \label{46} \end{equation} where $j\in Z$ and $-j\leq m,\bar m\leq j$. (\ref{46}) reduces to (\ref{42}) when $\lambda=1$. Here we see that we can adjust the extra modulus $\lambda$ in order to obtain Landau levels in the large volume limit. However the coefficient is not related to the magnetic field H, since the cancelations in the last term are due to a tuning of the modulus $\lambda$ which takes large (or small via duality) values. The interpretation of this limit is the following. Let us first take $H=0$ since it is not relevant in this limit. {}From the point of view of the coset space $SU(2)/U(1)$, the SU(2) WZW model can be viewed as a Dirac monopole on $S^2$, \cite{bk}. Thus at $\lambda=1$ we can write (\ref{46}) in the form \begin{equation} \Delta E_{j,m,\bar m}= {1\over R^2}\left[j(j+1)-m^2\right]+{m^2\over R^2}={j(j+1)\over R^2} \label{466}\end{equation} In (\ref{466}) the piece $j(j+1)-m^2$ of the energy is the standard spectrum of charged particles in the presence of the monopole and the additional $m^2$ is coming from the Kaluza-Klein masses of the charged modes. The states with non-trivial $\bar m$ are not charged with respect to the monopole and thus do not contribute to the energy. When we perturb $\lambda$ away from 1 we can suppress the Kaluza-Klein masses and thus we can have a limit similar to that of Eq. (\ref{36}). If we include all higher order corrections in $\alpha'$ and identify $R^2\to k+2$ then (\ref{46}) becomes \begin{equation} \Delta E_{j,m,\bar m}= {1\over k+2}\left[j(j+1)-m^2\right]+{\left(2\sqrt{k+2}eH-\left(\lambda+{1\over \lambda}\right)m-\left(\lambda-{1\over \lambda}\right)\sqrt{(1+2/k)}\bar m\right)^2\over 4(k+2)(1-2H^2)} \label{47} \end{equation} Eq. (\ref{47}) matches the string theory spectrum with the following identifications \begin{equation} m\to Q+J^3\;\;\;,\;\;\; e\to \sqrt{2\over k_{g}}\bar {\cal P} \;\;\;,\;\;\;\bar m\to \bar J^3 \end{equation} \begin{equation} H^2={1\over 2}\;\;{F^2\over F^2+2\left(1+\sqrt{1+F^2+{\cal R}^2}\right)} ={F^2\over 8}\left[1+{\cal O}(F^2,{\cal R}^2)\right] \end{equation} \begin{equation} \lambda^2= {1+\sqrt{1+F^2+{\cal R}^2}+{\cal R}\over 1+\sqrt{1+F^2+{\cal R}^2}-{\cal R}} =1+{\cal R}+ {\cal O}(F^2,{\cal R}^2) \end{equation} \vskip 1cm \section{Conclusions and Further Comments} We have studied a class of magnetic and gravitational backgrounds in closed superstrings and their associated instabilities. Our starting point are superstring ground states with an adjustable mass gap $\mu^2$ \cite{kk}. We have described in detail how to construct them starting from any four-dimensional ground state of the string, by giving appropriate expectation values to the graviton antisymmetric tensor and dilaton. In such ground states all gauge symmetries are spontaneously broken. Exact magnetic and gravitational solutions can then be constructed in such ground states as exactly marginal perturbations of the appropriate CFTs. In the magnetic case, there is a monopole-like magnetic field on $S^3$. The gravitational backreaction squashes mildly the $S^3$ keeping however an $SO(3)$ symmetry. We have calculated the exact spectrum as a function of the magnetic field. The first interesting observation is that, unlike field theory, there is a maximum value for the magnetic field $\sim M_{\rm planck}^2$. At this value the part of the spectrum that couples to the magnetic field becomes infinitely massive. We find magnetic instabilities in such a background. In particular, for $H\sim {\cal O}(\mu M_{\rm planck})$ there is a magnetic instability, driven by helicity-one states that become tachyonic. The critical magnetic field scales differently from the field theory result, due the different mechanism of gauge symmetry breaking. We also find that, unlike field theory, the theory becomes stable again for strong magnetic fields of the order $\sim {\cal O}(M^{2}_{\rm planck})$. Similar behavior is found for the gravitational perturbation. Here again there is an intermediate region of instability in the perturbing parameter. Such instabilities could be relevant in cosmological situations, or in black hole evaporation. In the cosmological context, there maybe solutions where one has time varying long range magnetic fields. If the time variation is adiabatic, then there might be a condensation which would screen and localize the magnetic fields. Results on such cosmological solutions will be reported elsewhere. Also, instabilities can be used as (on-shell) guides to find the correct vacuum of string theory. Our knowledge in that respect is limited since we do not have an exact description of all possible deformations of a ground state in string theory. Another subject of interest, where instabilities could be relevant is Hawking radiation. It is known in field theory that Hawking radiation has many common features with production of Schwinger pairs in the presence of a long range electric field. In open string theory it was found, \cite{bp} that this rate becomes infinite for a $finite$ electric field , $E_{\rm crit}\sim M^2_{\rm string}$ (unlike the field theory case) and this behavior is due to $\alpha'$ corrections. Notice also that in the open string it is $M_{\rm string}$ and not $M_{\rm planck}$ that is relevant due to the absence of gravity. It would be interesting to see if this behavior persists in the presence of gravity (which is absent to leading order in open strings) by studying the effect in closed strings. In fact we expect that gravitational effects will be important for $E\sim M^2_{\rm planck}$. For small $g_{\rm string}$ however, we can have $M_{\rm string} << M_{\rm planck}$ so we expect a similar behavior as in the case of open strings. It is plausible that similar higher order corrections modify the Hawking rate in such a way that (some) black holes are unstable in string theory. Such a calculation seems difficult to perform with today's technology but seems crucial to the understanding of stringy black holes. \centerline{\bf Acknowledgements} We would like to thank L. Alvarez-Gaum\'e, S. Coleman, M. Porrati and especially J. Russo for discussions. C. Kounnas was supported in part by EEC contracts SC1$^*$-0394C and SC1$^*$-CT92-0789.
\section{Introduction} The processes by which magnetization reversal occurs in the nanoscale ferromagnets that will make up the next-generation recording media are the subject of active research. One quantity for which theory and experiment often disagree is the lifetime $\tau$, which is the time required for a particle with initial magnetization $m_0 \! = \! +1$ to reach $m \! = \! 0$ when a magnetic field in the $-\hat{z}$ direction is applied. Micromagnetics,\cite{WFBrown} a theoretic technique in which differential equations are numerically solved on a coarse-grained lattice, predicts that the lifetime is given by the Arrhenius equation \begin{equation} \label{eq:Arrhenius} \tau \propto \exp (\beta \Delta F) \end {equation} with $\Delta F \! \propto \! L^d$. Here $\beta^{-1}$ is the temperature in units of energy, $\Delta F$ is the free-energy barrier between the stable and metastable phases, and $L$ is the linear system size. This same prediction is made by the standard N{\' e}el-Brown theory of single-domain ferromagnets.\cite{Neel49,Brown} The evident failure of Eq.~(\ref{eq:Arrhenius}) with $\Delta F \! \propto \! L^d$ for somewhat larger grains is ascribed to the existence of more than one domain in larger particles, as is a corresponding peak in plots {\em vs.}\ $L$ of the switching field $H_{\rm sw}$, which is the field required to yield a given lifetime. Recently techniques such as MFM have been used to resolve the magnetic properties of {\em isolated, well-characterized} single-domain particles (see, {\em e.g.}, Ref.~\cite{MFM}). This is an important advance, since previous experiments on ferromagnetic powders left uncertainties due to the range of grain sizes and orientations and the local magnetic environments. Observations of individual particles by MFM have made it clear that even for many {\em single-domain} particles, N{\' e}el-Brown theory is inadequate. We have applied the statistical-mechanical droplet theory of metastable decay to nanoscale ferromagnets with large uniaxial anisotropy, and compared Monte Carlo simulations of square-lattice Ising systems with droplet-theory predictions.\cite{2dpi} (For a review of droplet theory, see Ref.~\cite{RikARCP94}.) The agreement between theory and simulation is quite good, and despite the crudeness of the Ising model as a model for real magnets, it shows good qualitative agreement with the MFM experiments. We find rich $L$-dependent behavior in the standard Ising model, even though its equilibrium structure is a single domain for all $L$. This suggests that for some strongly anisotropic magnetic materials, magnetization reversal may occur through the nucleation and growth of non-equilibrium droplets. Details of our work are given in Refs.~\cite{2dpi} and \cite{demag}, including formulae for general dimensionality. For simplicity, we only discuss the two-dimensional case here. \section{Applied Droplet Theory} To be concrete, consider a two-dimensional kinetic Ising ferromagnet ($s_i \! = \! \pm 1$) with Hamiltonian \begin{equation} \label{eq:hamiltonian} {\cal H} = -J \sum_{\rm n.n.} s_i s_j - H \sum_i s_i + L^{-2} D \biggl( \sum_i s_i \biggr)^2 \end {equation} and Metropolis single-spin-flip dynamics on a square lattice with periodic boundary conditions. The last term represents a mean-field approximation for the dipole-dipole interaction energy and is taken to be zero except as noted. The critical radius of a ``droplet'' of $s_i \! = \! -1$ spins surrounded by $s_i \! = \! +1$ spins occurs when the free energy of the droplet ($2 \pi R \sigma \! - \! \pi R^2 2|H|$) is maximum: \( R_c \!\approx \! \sigma / 2|H|, \) where $\sigma$ is the surface tension per unit length. Droplets smaller than this will very probably shrink and vanish; larger droplets will very probably grow and reverse the magnetization of the system. In a sufficiently large system, the probability per unit time that a critical droplet forms, centered at a given site, is given by droplet theory as \cite{RikARCP94} \begin{equation} \label{eq:NucRate} I \propto |H|^3 \exp \left( - \beta \pi \sigma^2 / 2 |H| \right) \ . \end {equation} The details of the magnetization reversal depends on the number of critical droplets the system forms. \begin{figure} \vspace*{2.7truein} \special{psfile = hs_t182.psc angle = 90 hscale = 45 vscale = 45 hoffset = 330 voffset = -15} \caption{ The relation between the switching field $H_{\rm sw}$ and the system width $L$ for two different fixed lifetimes (solid curves), calculated by kinetic Ising model simulations at $k_{\rm B}T \! = \! 0.8 k_{\rm B}T_{\rm c} \! \approx \! 1.815 J.$ The dotted curve is near the crossover between the CE and SD regions;\protect\cite{2dpi} the dash-dotted curve is near the crossover between the SD and MD regions. \protect\cite{2dpi} } \label{fig:H_sw} \end{figure} For weak fields or small systems ($L \! < \! R_c$), no critical droplet can form. This is called the Coexistence (CE) Region, and since two interfaces (remember: periodic boundary conditions) must form to reverse the magnetization, \begin{equation} \label{eq:CELife} \tau \propto \exp \left\{ \beta \left[ 2 \sigma L - O\left(HL^2\right) \right] \right\} \ . \end {equation} For slightly larger $L$, the first supercritical droplet will grow to the size of the system before another one can form. The lifetime in this Single Droplet (SD) Region is \begin{equation} \label{eq:SDLife} \tau \approx \left[L^d I \right]^{-1} \ . \end {equation} In both the CE and SD regions, switching is a Poisson process, so the standard deviation of the lifetime is comparable to $\tau$. Both Eq.~(\ref{eq:CELife}) and Eq.~(\ref{eq:SDLife}) are actually special cases of the Arrhenius equation, Eq.~(\ref{eq:Arrhenius}), but in neither case is $\Delta F$ proportional to $L^d$. Note that if $\tau$ is held constant and the system size is increased, Eq.~(\ref{eq:CELife}) implies that the magnetic field must {\em increase}, whereas Eq.~(\ref{eq:SDLife}) implies that the magnetic field must {\em decrease}. This shows that the peak in $H_{\rm sw}$ occurs near the crossover between the CE and SD regions (see Fig.~\ref{fig:H_sw}). \begin{figure} \vspace*{2.7truein} \special{psfile = sd_p_h.psc angle = 90 hscale = 45 vscale = 45 hoffset = 330 voffset = -15} \caption{ The probability that $m \! > \! 0$ for a kinetic Ising system in the SD region. $T \! = \! 0.8 T_c$, $t \! = \! 914$ Monte Carlo steps per spin (MCSS) and $L \! = \! 10$. The solid curve is a fit of $\exp(-t/\tau)$ to the MC data, where $\tau$ is given by Eq.~(\protect\ref{eq:SDLife}). The inset figure shows the fitted curve over a wider range in $H$. } \label{fig:SD_P} \end{figure} \begin{figure} \vspace*{2.7truein} \special{psfile = md_p_h.psc angle = 90 hscale = 45 vscale = 45 hoffset = 330 voffset = -15} \caption{ The probability that $m \! > \! 0$ for a kinetic Ising system in the MD region. $T \! = \! 0.8 T_c$, $\tau \! = \! 40.7$ MCSS and $L \! = \! 30$, 100, and 300. The solid curves are fits of droplet theory predictions to the MC data. The dashed curve is the fit of the droplet-theory prediction for $L \! = \! 100$ extrapolated to $L \! = \! 300$. } \label{fig:MD_P} \end{figure} The probability that the magnetization is greater than zero $P(m \! > \!0)$ is shown as a function of field in Fig.~\ref{fig:SD_P} for a system in the SD region. This probability is what is most easily observed in MFM experiments, and it decays exponentially with time in both the CE and SD regions. In the SD region, the system is very unlikely to return to the the metastable state from the stable state, so \begin{equation} \label{eq:SD_P_t} P(m \! > \!0) = \exp \left(-t/\tau \right) \; . \end {equation} In the CE region such backwards switching takes place on a timescale comparable with the initial decay, so the situation is more complicated. For sufficiently large $L$ or $H$, several supercritical droplets may form before any one of them has grown to the size of the system. This is the Multi-Droplet Region (MD). Such systems were first studied by Kolmogorov,\cite{Kolmogorov37} Johnson and Mehl,\cite{JohnsMehl39} and Avrami,\cite{Avrami} and have a lifetime \begin{equation} \label{eq:MDLife} \tau \approx \left[ I\pi v^2/3 \ln 2 \right]^{-1/3} \ , \end {equation} where the radial growth velocity $v$ is assumed to be proportional to $|H|$. Although $\tau$ is independent of $L$, the variance of the lifetime is proportional to $(v/L)^2$. Measuring $P(m \! > \! 0)$ as a function of $H$ or $t$ thus provides a means of estimating the proportionality constant between $v$ and $H$ (see Fig.~\ref{fig:MD_P}). Details are given in Ref.~\cite{2dpi}. The addition of the dipole-dipole interaction energy in Eq.~(\ref{eq:hamiltonian}) makes the forms for the lifetime somewhat more complicated than we have presented here. In the CE and SD regions, a form of the Arrhenius equation [Eq.~(\ref{eq:Arrhenius})] still applies, and switching is still a Poisson process, so Eq.~(\ref{eq:SD_P_t}) still applies for the SD region. In the MD region, the system evolves in a time-dependent effective field, $H_{\rm eff} \! \equiv \! H - 2Dm(t)$. The time-dependent magnetization can be found to $O(D)$ fairly easily, and we find numerically that the $O(D^2)$ correction is relatively small. We can then solve analytically for $\tau$ to $O(D^2)$ and find good agreement with simulation results (Fig.~\ref{fig:md_tau_D}). A detailed treatment of the $D \! > \! 0$ case is given in Ref.~\cite{demag}. \begin{figure} \vspace*{2.7truein} \special{psfile = md_tau_D.psc angle = 90 hscale = 45 vscale = 45 hoffset = 330 voffset = -15} \caption{ The lifetime $\tau$ for a kinetic Ising system in the MD region, normalized by the lifetime in a similar system with $D \! = \! 0$. $T \! = \! 0.8 T_c$. The solid curves are droplet-theory predictions. } \label{fig:md_tau_D} \end{figure} \section{Conclusion} We have used Monte Carlo methods to simulate magnetization switching in two-dimensional kinetic Ising ferromagnets. The results of the simulations can be well explained by droplet theory and show good qualitative agreement with experiments, despite the comparative simplicity of the Ising model. This simplicity, in turn, allows us to develop an understanding of the underlying statistical mechanics. Particular features to make the model more realistic, such as appropriate boundary conditions, quenched randomness, and less rigorous anisotropy will be added in later studies. \acknowledgments We thank S.~von Moln{\' a}r, D.~M.\ Lind, J.~W.\ Harrell, and W.~D.\ Doyle for useful discussions. This research was supported in part by FSU-MARTECH, by FSU-SCR under DOE Contract No.\ DE-FC05-85ER25000, and by NSF Grants No.\ DMR-9315969 and DMR-9520325.
\section*{Figure Captions} \bigskip \noindent{\bf Figure 1:} The predicted light-element abundances (with $2\sigma$ theoretical errors); rectangles indicate consistency intervals, which all overlap for $\eta =(2.5-6)\times 10^{-10}$. \medskip \noindent{\bf Figure 2:} Likelihood functions for D and $^3$He (lower solid curves, from left to right: Models 1, 0, and 2), $^4$He (dotted curves, from left to right: $\sigma_Y=0.01$, $\sigma_Y=0.003$, and $\Delta Y= 0.01$), and $^7$Li (broken = high $^7$Li, solid = low $^7$Li). \medskip \noindent{\bf Figure 3:} Reduced $\chi^2$ as function of $\eta$ for eight different sets of assumptions about the systematic uncertainties (see text for details). \medskip \noindent{\bf Figure 4:} The 5\% of maximum likelihood contours for ${\cal L}(N_\nu, \Delta Y, \sigma_Y=0.003)$ (solid curves = low $^7$Li, broken curves = high $^7$Li). Because $N_\nu$ and $\Delta Y$ are not independent parameters, the contours of likelihood are diagonal lines and the likelihood function is not compact. \medskip \noindent{\bf Figure 5:} The likelihood function ${\cal L}(\sigma_Y, \eta ,\Delta Y=0)$ (solid curves = low $^7$Li, broken curves = high $^7$Li). \medskip \noindent{\bf Figure 6:} Same as Fig.~5 for ${\cal L}(\Delta Y, \eta , \sigma_Y = 0.003)$. \end{document}
\section{ I. Derivation of the ``First Law''} The work I will describe in this section, done together with Madhavan Varada\-rajan [3], grew out of our wish to understand what happens to the theorem that stationarity implies extremality, when spacetime has a boundary. It has been known for a long time that for gravity, or any other Lagrangian field theory, every solution of the field equations which has a Killing vector also has a corresponding extremality property: the conserved quantity associated to the Killing vector is unchanged by infinitesimal perturbations of the fields. Bernard Schutz and I had found a proof of this which we liked [1] and we wondered at the time what would happen if we applied it to a spacetime containing a black hole. The main message I want to leave you with today, is that what happens is that the so-called first law of black hole thermodynamics emerges in a very direct manner. The derivation which results in this way is of interest mainly because of its conceptual simplicity, but it also shows one new thing. It shows that the 3-surface $\Sigma$ on which the energy is evaluated can meet the black hole horizon anywhere; it doesn't have to go through any special place like a bifurcation submanifold\footnote{$^{1}$} {as it does, for example, in reference [4].}. I believe this is important, because the ability to push $\Sigma$ forward along the horizon is crucial to understanding where the {\it second} law of black hole thermodynamics comes from [5]. The proof also makes clear why the first variation of the energy gets contributions only from the horizon itself, and it provides an explanation (via the Raychaudhuri equation) of why it is specifically the change in horizon {\it area} which governs the change in the energy. The derivation also illustrates how integral formulations of conservation laws can often be more convenient than differential ones. It takes place in 4D for Einstein gravity (with a possible electro-magnetic field), but there is no reason it could not be extended to higher dimensions, or to other lagrangians. The proof is also in such a form that it might help in understanding the behavior of the {\it second} variation of the energy. This variation is important in connection with stability, but I don't have any new results to report on it. Since a detailed account will soon be available [3], there is no reason to try for completeness here. Instead I will present the main steps of the analysis as simply as I can, in a manner which I hope will be complementary to that of reference [3]. In the same minimalist spirit I will mainly restrict myself to the case of pure Einstein gravity and will set the electromagnetic field and black hole rotation rate to zero. \subsection{The Noether operator and the total energy} Before we can get to the proof proper, we need the notion of Noether operator and a technique I will call asymptotic patching. The Noether operator formalizes her explanation of how continuous symmetries imply conservation laws. For a first-order Action $S$ depending on dynamical fields $Q$ and background fields $B$, and for a geometrical symmetry like energy or angular momentum, the Noether operator is defined through the identity, $$ \delta_{(f,\xi)} S = \oint_{\partial{\cal X}} f \, {\cal T}^a_b\cdot\xi^b d\sigma_a - \int_{\cal X} {\delta {\cal L} \over \delta Q} \, f \Lie_\xi Q \, dV \eqno(1) $$ Here the variation $\delta_{(f,\xi)}$ is what might be called a ``partial dragging'' of both the fields $Q$ and the region of integration ${\cal X}$ through the vectorfield $\xi$, specifically it drags $\partial{\cal X}$ by an amount $f\xi$ and it alters $Q$ by $\delta_{(f,\xi)}Q = - f\Lie_\xi Q$. If there are no background fields $B$ present in the lagrangian ${\cal L}={\cal L}[Q;B]$ --- or if $\xi$ is a symmetry of those which are present --- then for $f\equiv 1$, the variation $\delta_\xi{S}$ evidently vanishes identically. Now the total energy of a solution, evaluated on a surface $\Sigma$ which is the future boundary of a spacetime region ${\cal X}$, can be defined as the value of $\delta S({\cal X})$ when $\Sigma$ and all the fields on it are translated in time in such a manner as to hold fixed the boundary conditions at infinity [6]. Choosing $f$ and $\xi$ to implement these requirements and assuming sufficiently rapid falloff of $Q$ at infinity leads directly to the formula for the energy $$ - E = \int_\Sigma {\cal T}^a_b\cdot\xi^b \, d\sigma_a, \eqno(2) $$ where $\xi$ is any vectorfield which is a (future-directed) time translation in a neighborhood of infinity (an exact Killing vector of the flat background there.) In a situation where the background either is absent or enters only as a surface term in the Action, a compensating deformation by $-f\xi$ reduces $\delta{S}$, and therefore $E$, to a surface integral at spatial infinity. In virtue of (2), this has the formal consequence that ${\cal T}^a_b\cdot\xi^b$ must take the form of a pure divergence $\partial_b({\cal W}^{ab}\cdot\xi^b)$ plus a term which vanishes ``on shell''. Specializing to gravity (and for brevity omitting indices and indications of elements of surface/volume and of density-weight), we have specifically $$ T \cdot \xi = {\rm div} (W\cdot\xi) - G \xi, \eqno(3) $$ so that (2) reduces on shell to\footnote{$^{2}$} {For angular momentum, everything would be the same, except that $\xi$ would be a rotation generator near spatial infinity, instead of a time translation. The explicit form of $W$ can be found in refs. [6], [7] and [3].} $$ - E = \oint_\infty W \, . \eqno(4) $$ \subsection{Asymptotic patching} In computing the energy, etcetera, of an asymptotically flat solution $g_{ab}$ to the Einstein equation, we don't directly use the covariant Action ${1\over{2}}\int{RdV}$. Instead we do something else which can be described in different ways. Perhaps the best description is just that we replace the metric $g$ by one which is strictly flat near spatial infinity. In reality the very long range part of the metric representing an isolated system like a black hole is meaningless in any case, since it cannot be isolated from the fields of other objects which are invariably present. Hence, there should be no distinction, in a physical sense, between $g_{ab}$ and a metric which has been ``cut off'' at large radii by ``patching'' it to a flat metric. The $S$ whose variation yields the energy of the isolated system is best viewed, I believe, as nothing but the covariant Action\footnote{$^{3}$} {For many purposes, one also needs to add to this Action a surface term like ${\rm tr}{K}$ integrated over the initial and final spacelike boundaries; but here we can ignore any such addition since it does not contribute to the variation defining $E$.} of this cut off field $\widetilde g_{ab}$; and the technique for obtaining $\widetilde g_{ab}$ by gradually deforming the original metric to the flat one as some radial parameter $r$ increases from $R$ to $2R$ is what I mean by ``asymptotic patching'' [1][3]. Asymptotic patching can also be understood in a purely technical way in relation to an integration by parts performed to render the Action finite. For generic O($1/r$) falloff in the metric, the Ricci scalar $R$ will decay only like $1/r^3$, which leads to a logarithmically divergent integral for $S$. By adding a suitable divergence to the integrand we eliminate from $R$ the terms of the form $g\partial\ptl{g}$, thereby improving its falloff to $1/r^4$, while at the same time making $S$ first-order so that the above definition of the Noether operator applies without modfication. The improved falloff suffices to render both $S$ and $E$ well-defined.\footnote{$^{4}$}% {For angular momentum a strengthened asymptotic condition is needed (``parity condition'').} Having modified $S$ in this way, we can then perform the same patching to a flat metric at infinity without producing any further change in the Action or the energy (in the limit in which the patching radius $R$ recedes to infinity) [1][3]. This second viewpoint is perhaps somewhat more advantageous technically, but it requires the introduction of extra background: a globally defined connection with respect to which one can perform the integration by parts. The upshot from either point of view is that we end up having to deal only with metrics which are strictly flat near infinity. This will free us from having to worry about the effects of variations at infinity, leaving only boundary terms at the horizon to contribute. It also means, of course, that we can no longer express the energy as the flux integral (4) taken at true infinity, but (4) still holds if evaluated {\it just inside} the patching radius, and the expression (2) in terms of a spatial integral remains generally valid, under the assumption (which will always be in force) that $\xi^a$ remains an exact Killing vector of the flat asymptotic metric throughout the patching region. Henceforth, we will consider only metrics which have been patched to become strictly flat near $\infty$, and "solution" will always mean solution patched to flat metric at large $r$. In addition, we will consider only vectorfields $\xi$ which preserve any background which has been introduced, and which in particular are strict Killing vectors (of the flat metric) near infinity. \subsection{The extremum proof without reference to a horizon} Setting $f=1$ in the defining identity (1) of the Noether operator, and recalling that the left hand side then vanishes automatically, we obtain the basic identity $$ \oint T\cdot\xi = \int {\delta L\over\delta g} \Lie_\xi g , \eqno(5) $$ for an arbitrary metric $g$ and vectorfield $\xi$. (Here $\delta L / \delta g$, if made explicit, would be $-G^{ab}$ of course.) Now consider (Figure 1) a spacetime region ${\cal X}$ bounded to the past and future by asymptotically flat surfaces $\Sigma_0$ and $\Sigma$, and let the metric $g$ be {\it a stationary solution} to the Einstein equation. \epsfxsize=3.6in \FigureNumberCaption{fig-1.eps}% {1}% {The spacetime region ${\cal X}$ involved in proving energy extremality without reference to a horizon. It is bounded to the future and past by the surfaces $\Sigma$ and $\Sigma_0$.} \noindent If $g'$ is a nearby solution (not necessarily stationary) then it is easy to cobble together a perturbation $\delta g$ which vanishes in a neighborhood of $\Sigma_0$ and for which $\delta g = g'-g \;$ in a neighborhood of $\Sigma$. Let us apply the identity (5) to this perturbation, in fact let us consider the result of perturbing $g$ in (5) by an arbitrary $\delta g$. On the RHS we have the product of two expressions which both vanish for the unperturbed metric; the product therefore remains zero to first order in the perturbation; consequently the LHS must also vanish, i.e. $$ \delta \oint_{\cal X} T\cdot\xi = 0 \eqno(6) $$ {\it for an arbitrary perturbation $\delta g$ and an arbitrary region ${\cal X}$}. But for our $\delta g$ this expression itself is by (2) none other than the difference $E(g)|_{\Sigma_0} - E(g')|_\Sigma$, which accordingly must vanish. In other words $E'=E$ or $\delta E = 0$, where now $\delta E$ just means the variation in $E$ on $\Sigma$ in going from $g$ to $g'$. This is our first main result: the total energy is an extremum for any asymptotically flat stationary solution to the field equations. \subsection{Application to a spacetime with internal boundary} Thus far I have been tacitly assuming that the 3-surface $\Sigma$ is a complete Riemannian manifold possessing a sole asymptotic region. When spacetime has more than one asymptotic region, or more importantly for us, when it has an {\it internal boundary}, the formula (2) for $E$ must be applied with care. In order that it correctly furnish the energy associated with a given $\infty$, $\xi$ must be a time translation there, but it must vanish at all the other boundaries (including the actual internal ones and the ideal ones at infinity). This rule follows from the prescription that $E$ represents the change in $S$ which results from a perturbation that {\it rigidly displaces the entire spacetime relative to the infinity} in question. Alternatively, it can be derived by reverting to the formula (4) for the energy of a ``non-patched'' solution, and converting (4) to a volume integral via Stokes theorem. In the case of interest the boundary will be the horizon of a black hole (or holes). This surface does not represent a physically real ``edge'' of spacetime, of course, but rather a boundary we impose on the submanifold we work with, in order to make effective use of the identity (5). Being a future horizon, this bounding surface (which I will call $H$) will be null with its future side facing away from the outer world.\footnote{$^{5}$} {It is instructive to examine the reasons why the theorem just proved for spacetimes without boundary does not furnish useful information when black holes are present. In the maximally extended Schwarzschild spacetime, for example, the theorem does apply, but, since there are two infinities the extremized energy $E$ is the {\it sum} of the masses seen from the infinities; and this in turn vanishes since the requirement that $\xi$ be Killing forces it to point backward in one of the asymptotic regions. Thus the theorem is indeed obeyed, but yields only the trivial fact that $\delta(zero) = zero$ ! To make $E$ be the physically relevant energy, we could eliminate the second infinity via an antipodal identification (leading to a geon spacetime with spatial topology ${{\rm I\!\rm R}}P^2\times{\rm I\!\rm R}^3$), but then one would encounter an inconsistency in trying to extend $\xi$ inside the horizon as a Killing vector: the identified spacetime would no longer be stationary in the sense required by the theorem. Either way, we fail to gain useful information by trying to apply the theorem to the manifold as a whole.} Let us now try to generalize the reasoning of the previous subsection to a region ${\cal X}$ formed as before (with future boundary $\Sigma$ and past boundary $\Sigma_0$) but with an extra internal boundary $H$ representing the portion of the horizon between $\Sigma_0$ and $\Sigma$. In referring to this setup I will denote the 2-surface $\Sigma\cap H$ by $S$, and the corresponding, but earlier, 2-surface $\Sigma_0\cap H$ by $O$. (See Figure 2.) \epsfxsize=3.6in \FigureNumberCaption{fig-2.eps}{2}% {A region ${\cal X}$ analogous to that of Figure 1, but truncated at the horizon. The null surface $H$ is that portion of the horizon between $\Sigma_0$ and $\Sigma$; its future boundary is the 2-surface $S$ and its past boundary $O$.} Now in order to use the identity (6) as we did in the previous subsection, we need $\xi$ to be a Killing vector of the unperturbed solution, which contradicts the requirement that it vanish on $H$ in order that (2) be the total energy. However $\xi$ {\it is} a Killing vector at large radii, so there is nothing to stop us from making it Killing everywhere by use of the relation (3). Applying this relation in conjunction with Stokes' Theorem to the region $\Xi\subseteq\Sigma$ where $\xi$ deviates from being Killing, we immediately obtain\footnote{$^{6}$} {by converting the integral of $T\cdot\xi$ over $\Xi$ to an integral over $\partial \; \Xi$ of $W\cdot\xi$, then making $\xi$ Killing within $\Xi$, then converting back to an integral over $\Xi$.} $$ - E = \int_\Sigma T\cdot\xi + \oint_{S} W\cdot\xi, \eqno(7) $$ where now $\xi$ is Killing everywhere and the integral over $S$ appears because $S$ is the inner boundary of the region $\Xi$. Expressed in this manner, the energy appears as a volume integral augmented by a horizon contribution which it would be natural to desribe as the ``energy of the black hole''. Now let us apply to (7) a variation leading from the stationary solution $g$ to a nearby solution $g'$, and let us temporarily assume that $g'=g$ in a neighborhood of $S$. Since the variation of the second integral in (7) then vanishes trivially, exactly the same proof as earlier shows that $\delta E=0$. From this it follows immediately that $\delta E$ for a general perturbation {\it can depend only on the value of the perturbation (and its derivatives) at the horizon itself}, i.e. at the 2-surface $S$. Notice that essentially no computation was involved in reaching this conclusion. Consider, then, a perturbed solution $g'$ for which $g'-g$ does not necessarily vanish on the horizon. We can study its energy by introducing the same kind of ``interpolating perturbation'' $\delta g$ as we used earlier in the absence of a boundary; however before doing this, it will be convenient to prepare ourselves by extending the $\Sigma$-integral in (7) all the way back to $\Sigma_0$ with the aid of the identity (3). The result is $$ - E = \int_{\Sigma\cup H} T\cdot\xi + \int_H G\xi + \int_O W\cdot\xi . \eqno(8) $$ Now when we apply the variation $\delta$, the first integral in (8) is unchanged for exactly the same reason as earlier and we are left with $$ - \, \delta E = \delta \int_H G \; \xi \eqno(9) $$ (the third integral in (8) being trivially unchanged because $\delta g $ vanishes in a neighborhood of $\Sigma_0$). This is our second main result. It expresses $\delta E$ as the change in the flux of the fictitious (conserved) energy current $G^a_b\xi^b$ across the horizon $H$ in going from the stationary to the varied solution.\footnote{$^{7}$} {For the case of a rotating black hole, the relevant Killing vector would be $\xi=t+\Omega\phi$ where $t$, now, denotes the Killing vector which is a time-translation at infinity, while $\phi$ denotes the rotational one ($\Omega$ being the angular velocity of the horizon). Hence the variation $\delta E$ in (9) would be replaced by $\delta E - \Omega\delta J$, $J$ being the angular momentum.} Notice that all reference to auxiliary background fields has now dropped out. \subsection{Reduction of $\delta E$ to an integral on S } We have already seen on general grounds that $\delta E$ must be expressible in terms of quantities defined only on the 2-surface $S$ in which our 3-surface $\Sigma$ meets the horizon. To discover the explicit form of this expression requires us to convert (9) from an integral over $H$ to one over $S$ alone. Clearly, it suffices to re-express it as the integral of a total divergence of some ``potential''.\footnote{$^{8}$} {Another approach would be to shrink $H$ to a 2-surface by bringing together the surfaces $\Sigma_0$ and $\Sigma$.} It turns out that there is a systematic method [8] for constructing such a potential (and the potential is uniquely determined by the construction); its applicability is guaranteed by the fact that $\delta(\sqrt{-g}G^a_b\xi^b)$ vanishes for arbitrary variations $\delta{g}$.\footnote{$^{9}$} {This identity (cf. ref. [7]) is the analog of eq. (6) for the covariant Action ${1\over{2}}\int{RdV}$, only expressed in differential form. It can be derived as such, but it also follows immediately from eqs. (6) and (3).} To apply this method would require only straightforward calculation, but instead of following this route, we can invoke the Raychaudhuri equation to evaluate the integral in (9) directly, an approach which --- though it is less systematic than the method of reference [8] --- affords a simple explanation of how the horizon area emerges as the measure of $\delta E$. In essence, all that is involved is using the Raychaudhuri equation to convert the integrand of (9) into an expression involving the expansion of the horizon, and then performing an obvious integration by parts. In preparation, however, we need to recall a few definitions and make a convenient choice of gauge in which to express the perturbation $\delta{g}$. Let us begin by noting that for a non-rotating, stationary black hole (Schwarz\-schild metric), the timelike Killing vector $\xi$ is automatically null on the horizon,\footnote{$^{10}$} {The need for $\xi$ to be null is what forces us to take $\xi=t+\Omega\phi$ in the rotating case, as described in an earlier footnote.} whence proportional to the null geodesic generators of $H$. Accordingly, if we parameterize the latter with an affine parameter $\lambda$, then we have $$ \xi^a = \alpha {dx^a \over d\lambda} \eqno(10a) $$ for some function $\alpha$ depending on the choice of normalization for $\lambda$. Now although $\alpha$ is not uniquely determined, its $\lambda$-derivative is, and is given by $$ {d\alpha \over d\lambda} = \kappa, \eqno(10b) $$ where the black hole's ``surface gravity'' $\kappa$ is defined by the equation $\xi^b\nabla_b\xi^a=\kappa\xi^a$. Now in comparing the stationary solution $g$ with the interpolating metric $g+\delta g$, we are free to choose the diffeomorphism-gauge so that the horizon is the same surface $H$ for both metrics. In fact we clearly can arrange that $\xi$ remains a null generator of $H$ with respect to $g+\delta{g}$ and also that $\lambda$ remains an affine parameter along every such generator. (For given choices of $\Sigma_0$ and $\delta g$, this also determines to first order in $\delta g$ where $\Sigma$ is embedded in the varied spacetime.) In this gauge, equations (10a,b) will remain true even after the variation (with $\kappa$ denoting the surface gravity of the {\it unvaried} metric $g$, as always.) Finally we will use the fact that the extensor\footnote{$^{11}$} {I propose to call ``extensors'', the various tensorial objects which represent infinitesimal portions of submanifolds in expressions denoting integrals over such manifolds, for example $d\sigma_a$ and $dV$ in eqs. (1) and (2), or the codimension-two extensor $dS_{ab}$ which is implicit in the second integral of eq. (7).} $dS_a$ representing an element of the surface $H$ can be written as $$ dS_a = - \dtA \, dx^a \eqno(11) $$ for a portion of $H$ with cross-sectional area $\dtA$ and extension along the null direction in $H$ given by the (future pointing) null vector $dx^a$. Now we are ready to substitute into (9) the Raychaudhuri equation for the generators of $H$, namely $$ R_{ab}{dx^a\over d\lambda}{dx^b\over d\lambda} = - {d\theta\over d\lambda} - {\theta^2\over 2} - {\sigma^2\over 2}. \eqno(12) $$ Since $\theta=\sigma=0$ in the unvaried spacetime ($\theta$ = expansion, $\sigma$ = shear), only the first term on the RHS of (12) survives variation; and since $R_{ab}$ vanishes for the unvaried solution as well, we can write $R_{ab}=\delta R_{ab}$ and $\theta=\delta\theta$. Using these facts, we can replace (12) for the metric $g+\delta g$ with the equation $$ R_{ab}{dx^a\over d\lambda}{dx^b\over d\lambda} = - {d\theta\over d\lambda}. \eqno(13) $$ Noting further that $G^{ab}$ also vanishes for the unvaried metric, we can now transform (9) as follows: $$ \eqalignno { - \delta E &= \delta \int_H dS_a G^a_b \xi^b \cr &= \int_H dS_a G^a_b \xi^b \cr &= - \int \dtA (dx)_a G^a_b \xi^b \qquad ({\rm by\ eq.\, (11)}) \cr &= - \int \dtA\,(dx)_a G^a_b \alpha {dx^b\over d\lambda} \qquad ({\rm by\ eq.\, (10a)}) \cr &= - \int \dtA\,d\lambda\;\alpha\; G_{ab}{dx^a\over d\lambda} {dx^b\over d\lambda} \cr &= - \int \dtA\,d\lambda\;\alpha\; R_{ab}{dx^a\over d\lambda} {dx^b\over d\lambda} \cr &= \int \dtA\,d\lambda\;\alpha\,{d\theta\over d\lambda} \qquad ({\rm by\ eq.\,(13)}) \cr &= - \int_H \dtA\,d\lambda\; {d\alpha\over d\lambda}\, \theta + \int_S \dtA\;\alpha\;\theta \cr &= - \int_H d\lambda {d (\dtA)\over d\lambda} \kappa + \int_S \dtA \alpha \theta \cr &= - \int \kappa \int d\lambda {d (\dtA)\over d\lambda} + \int_S \dtA \alpha \theta \cr &= - \int_S \kappa \delta (\dtA) + \int_S \dtA \alpha \theta \cr &= - \kappa \delta A + \int_S \dtA \alpha \delta\theta \cr } $$ Here in the sixth equality we used that $dx/d\lambda$ is null; in the ninth we used eq. (10b) and that $\theta$ can be expressed as $$ \theta = {1 \over \dtA} {d(\dtA)\over d\lambda} ; $$ and in the last equality we used that $\kappa$ is constant on $H$. Thus we have reduced to $\delta E$ to an expression pertaining solely to the cross section $S$ of the horizon, and this is our third main result: $$ - \, \delta E = -\kappa \delta A + \int_S \dtA\,\alpha\,\delta\theta \eqno(14) $$ \subsection{Locating the horizon} With equation (14) our work is essentially complete, except that, in addition to the desired (first) term it contains an integral depending on $\delta\theta$. In order to realize why this unwanted term is present, we only have to ask ourselves where we have used the fact that $H$ is actually the horizon of the perturbed solution $g'$, and not just some random null surface therein. The point is, of course, that we haven't used it yet, meaning that the area of $S$ might have changed just because it was displaced in location without even leaving the unvaried spacetime! In order to distinguish such a bogus $\delta A$ from the true one, we need a criterion to locate the horizon with respect to the metric $g'$. Such a criterion, I claim, is precisely the requirement that $\delta\theta=0$ everywhere on $H$ (where here and henceforth `$\delta g$' just means $g'-g$, not the more complicated interpolating perturbation of earlier subsections). In principle this claim, if true, should be derivable from the Einstein equation, and such a derivation does not look too impractical, at least in connection with the Schwarzschild metric, whose perturbations are fairly well understood. For now however, we will derive $\delta\theta=0$ from an assumption which is a special case of the so-called cosmic censorship conjecture. We will {\it assume} that the horizon of the stationary solution $g$ cannot be destroyed by arbitrarily small perturbations of the metric. If this is so (and if it is not, then black holes do not exist in reality anyway!) then no infinitesimal perturbation of the metric $g$ can make the expansion $\theta$ negative anywhere, because if it did, then there would be arbitrarily nearby solutions $g'$ with negative expansion somewhere on their horizons, but it is well-known that negative expansion implies that the horizon encounters a singularity in a finite ``time'' (really affine parameter).\footnote{$^{12}$} {The argument uses the Raychaudhuri equation: positive convergence implies infinite convergence in a finite time, implies a generator leaves the horizon, implies a singularity.} But if $\theta=\delta\theta$ can never be negative, then it can never be positive either, because a simple change in the sign of $\delta g$ will similarly change the sign of $\delta\theta$ (and of course, $-\delta g$ will also be a solution of the linearized Einstein equation). Hence $\theta$ on the true horizon must remain zero to first order in any perturbation about a stationary black hole metric. \subsection{Summary: the first law} Our analysis is now complete; let us summarize the highlights. Using the identity (6) we first found that $\delta E=0$ for any variation $\delta g$ supported away from a cross-section $S$ of the horizon $H$. This implied that for general perturbations, $\delta{E}$ can depend only on the behavior of $\delta g$ in the neighborhood of $S$. To evaluate $\delta E$ explicitly, we had to re-express the 3-dimensional integral (9) as the integral of a divergence. A systematic method for doing so exists, but we used the Raychaudhuri equation instead, leading to equation (14). By invoking the ``stability'' of the horizon we ``situated'' $H$ within the perturbed spacetime, showing thereby that the second term in (14) is in fact zero when evaluated on the correctly identified perturbed horizon. The remaining term yields\footnote{$^{13}$} {For a rotating black hole, the substitution $\xi=t+\Omega\phi$ of earlier footnotes leads immediately to $-\delta E = -\kappa \delta A +\Omega \delta J$. For the charged case, a bit of extra analysis is needed, but again only general features of the theory are used, without any reference to the explicit form of the Kerr-Newman metrics (see [3]).} the so-called ``first law'' $$ \delta E = \kappa \delta A. \eqno(15) $$ \subsection{Possible further work} With $A$ identified as entropy, the fact that $\delta A=0$ whenever $\delta E=0$ can be interpreted as the first-order expression of thermodynamic stability (in the $\hbar\rightarrow 0$ limit), a thermodynamically stable solution being one which maximizes entropy at fixed energy. At second order, this maximization is generically equivalent to $$ A'' - \kappa^{-1} E'' \ge 0\ \ {\rm on}\ \ \ker E', \eqno(16) $$ where $(\cdot)'$ denotes Fr\'echet derivative. An interesting problem would be to try to prove (16) for Schwarzschild (say) by extending the foregoing analysis to second order.\footnote{$^{14}$} {In this connection, there might be extra conceptual complications in the rotating case, associated with the presence of so-called super-radiant modes.} Another worthwhile extension of the analysis would be to generalized gravity theories, including in a Kaluza-Klein setting (cf. [9]). There our ``Raychaudhuri trick'' would probably fail, and one would have to find another trick or fall back on the general method referred to earlier. Indeed this general method [8] merits following up even in ordinary gravity, both as a ``warmup'' for more complicated Lagrangians, and for the additional insight it might offer into the origins of the first law itself. \section{ II. Fractality of the Horizon } One possible source for the entropy of a black hole is in the fluctuations of a quantum field propagating near the horizon. When the field in question is the linearized metric (``graviton''), the associated entropy is geometrical in character, but there are many other quantum fields which are able to contribute as well. The mechanism in all cases is the same: fluctuations in the field occur on all scales, and when a fluctuation with characteristic size $\lambda$ is astride the horizon it sets up a correlation (``entanglement'') between inside and outside which metamorphoses into entropy when one ``traces out'' the field modes inside the black hole in order to obtain the effective density-operator describing the field outside the black hole [10]. When one tries to compute the value of this entropy for a free field, one obtains, at first, an infinite result deriving from the fact that free fields are scale-invariant in the ultraviolet regime, whence an infinite number of modes contribute with constant entropy per mode. However, if one introduces a cutoff at some scale $l$, the entropy takes on the finite value $S=c A/l^2$, where $A$ is the area of the horizon, and $c$ is a dimensionless constant of order unity. Since this gives the right area dependence, and also the right general magnitude if one chooses $l=l_{Planck}$, one is tempted to conclude on the one hand that one has explained black hole entropy, and on the other hand that one has obtained persuasive evidence for the existence of spatio-temporal discreteness in nature. Another thing which speaks in favor of identifying $S$ with some sort of entanglement entropy is that the prospect of a natural proof of the Second Law then arises naturally. Indeed, one can argue that, if full quantum gravity furnishes us (at some level of coarse-graining) with a well-defined, autonomously evolving density-operator $\rho$ describing the outside world, then $-{\rm tr}\rho\ln\rho$ necessarily increases as the surface $\Sigma$ with which it is associated moves forward in time. The argument [5] rests on the fact that the total energy is conserved and determinable from the gravitational field outside the black hole(s), no matter what may be occuring inside of the horizon (i.e. it rests on equation (4) or (7) above. Notice that the entropy does not change if the codimension-two surface $S$ in which $\Sigma$ meets the horizon does not move forward along $H$; hence the significance, referred to earlier, of being able to choose $S$ freely.) Although the argument just alluded to does not care what degrees of freedom it deals in (as long as the number of effective external states is finite at finite total energy), our interest here is in those variables associated with the fluctuations of quantum fields. To take seriously their contribution to the entropy leads to the seeming difficulty that --- for fixed discreteness scale $l$ --- the magnitude of $S$ would depend on the total number of fields present in nature, seemingly at odds with the simple geometrical character of the formula $S=2\pi A$, which just equates the entropy to the circumference of the unit circle times the area of the horizon measured in Planck units. (We take $l_{Planck}=\sqrt{\kappa}$ where $\kappa=8\pi G$ is the ``rationalized gravitational constant'', and $\hbar=c=1$.) This simple formula seems more in harmony with a directly ``geometrical'' character for the relevant degrees of freedom, perhaps the shape of the horizon itself [11], or the configuration of some underlying discrete structure composing the horizon, such as (the appropriate portion of) a causal set. The ``fractal'' picture of the horizon I will describe in a moment grew out of my wondering whether one could avoid the above ``species dependence problem'' by somehow writing the quantum fields out of the script in favor of more suitably geometrical degrees of freedom. In the meantime it has become much less clear that there is in fact any difficulty to be avoided, in view of the observation [12] that a change in the number of fields would affect not only $S$ but also the renormalized value of $\kappa\equiv 8\pi G$, and indeed would alter $\kappa$ in just the manner needed to compensate for the change in the entanglement entropy, leaving the formula $S=2\pi{}A/\kappa$ still valid. Although the details of their argument can be criticized, its overall structure is ``too pretty to be wrong'', and so is probably correct at some level. At the same time, it manifestly ignores the influence of the fluctuations on the horizon itself (``back reaction''), and to that extent is limited to a semiclassical regime. In the picture I am proposing, the number of species is irrelevant for an entirely different reason, namely for the reason that --- due precisely to the back-reaction --- the constant $c$ is not constant at all, but rather depends on the size of the black hole in such a manner as to become negligibly small for all but Planck sized black holes. More accurately I will try to show that the approximation of fixed horizon location and shape becomes invalid at a length scale much greater than Planckian, namely at a scale of the magnitude $M^{1/3}$, $M$ being the mass of the black hole.\footnote{$^{15}$} {In this and all subsequent formulas, we adopt units such that $8\pi{}G\equiv\kappa=1$.} Below that scale, the field fluctuations become strongly coupled to the horizon shape, and a semi-classical analysis becomes unreliable. At the same time, the shape of the horizon itself becomes ``fractal'' due to the effects of the fluctuations, perhaps providing the anticipated geometrical degrees of freedom to ``absorb'' the field ones. The point is that, at least for free or asymptotically free fields, fluctuations occur with equal intensity at all sufficiently small scales $\lambda$. Given a fluctuation of size $\lambda$, one would expect the associated energy of magnitude $\sim 1/\lambda$ to induce a concomitant distortion of the horizon. Heuristically, we may perhaps picture the situation as follows. As one descends in scale, one will reach a threshold size $\lambda_0$, at which the ``virtual energy'' of a typical fluctuation will be big enough to distort the horizon shape by an amount comparable to the size of the fluctuation itself. Then, like a sleeper who is uncomfortable in bed and either buries him/herself under the blankets or pushes them all on the floor, the fluctuation will either pull the horizon up over itself or (in the case of negative energy-density) drive the horizon entirely away. In either case the fluctuation will no longer overlap the horizon, and it therefore will no longer contribute to the entanglement entropy. Moreover, this effect evidently entails a strong coupling between the horizon shape and the field fluctuations of size $\lambda{\lower4pt\hbox{$\buildrel<\over\sim$}}\lambda_0$; whence such a fluctuations should not count as independently ``entangled'' degrees of freedom even if they do happen to meet the horizon. We conclude then, that the scale $\lambda_0$ sets a limit to our understanding of entanglement entropy, and that the only reliable estimates we can make for the latter pertain to fluctuations with characteristic sizes greater than $\lambda_0$. But isn't it obvious that $\lambda_0$ will just turn out to be of Planckian size in any case? To begin to answer this question reliably, one would have to analyze the effect on the horizon of a spacetime ``energy flux loop'' of characteristic size $\lambda$, situated on or near the horizon of, say, a Schwarzschild black hole. Here, I will do something less accurate but much easier: I will compute for {\it Newtonian} gravity, the disturbance in the ``horizon'' induced by a small additional mass $m\sim 1/\lambda$ distributed throughout a spatial region of size $\lambda$ located in the vicinity of the horizon. So let there be present at the origin a spherical mass M, and define its {\it horizon} as the locus of points where the escape velocity equals unity (i.e. $c^2$), that is, where $$ V = - 1/2, \eqno(17) $$ $V$ being the gravitational potential, $- GM/r$, of the mass.\footnote{$^{16}$} {One should presumably conceive of the perturbation as enduring only for a time of order $\lambda$, but the associated retardation effects would be hard to incorporate in the Newtonian framework, and in any case, they would not seem likely to alter the qualitative picture derived from treating the horizon as determined by the instantaneous Newtonian potential.} It will be convenient to work, not with $M$, but with the corresponding ``geometrized mass'' or ``Schwarzschild radius'' $R := 2GM$. In terms of $R$ we have for a point mass, $V(r)={}-{}R/2r$, so that the horizon occurs precisely at $r=R$, a well-known coincidence. Now let us add in the gravitational potential of the fluctuation, to which for analytical convenience, we will assign the effective mass-density $\rho=a\lambda/r_1(r_1+\lambda)^3$, resulting in the potential $$ V_1 = {-a/2 \over r_1 + \lambda}. $$ Here, $a$ is the net geometrized mass of the fluctuation, and $r'$ the distance to its center. Making the substitution $a=f/\lambda$ ($f$ being some fluctuation-dependent ``fudge factor'' of order unity) yields finally for the combined Newtonian potential $V$, $$ - 2V = {R\over r} + {f/\lambda \over \lambda + r' } $$ For simplicity, let us now place the center of the fluctuation where the unperturbed horizon meets the $y$-axis, and let us also move the origin of our coordinate system to that point. Then if we restrict ourselves to the positive $y$-axis, the potential assumes the particularly simple form $$ - 2V = {R\over y+R} + {f/\lambda \over \lambda + y}. \eqno(18) $$ In the approximation that $y,\lambda \ll R$ this reduces to $$ - 2V \approx 1 - {y\over R} + {f/\lambda \over \lambda + y}. \eqno(19) $$ It is now easy to locate the perturbed horizon by solving the equation $V=-1/2$ or $-2V=1$. Working with the approximation (19), we have on the horizon, $$ {y\over\lambda}({y\over\lambda}+1)={fR\over\lambda^3}, \eqno(20) $$ so that, if we denote by $h=y$ the height of the bulge raised in the horizon by the fluctuation, we see immediately that the relative height $h/\lambda$ depends only on the characteristic combination of parameters $fR/\lambda^3$. Moreover it is clear that $h/\lambda$ is of order unity when $fR/\lambda^3$ is, and that it becomes small for $fR/\lambda^3\ll{}1$. In other words, the threshold scale we are looking for is in fact $$ \lambda_0 \sim R^{1/3} $$ (where I have omitted $f$ since it is in any case of order unity). The width of the bulge can be determined similarly. Indeed, one finds after a little algebra that the profile of the bulge is determined by the equation, $$ {y\over\lambda} \left( 1 + \sqrt { ({x\over\lambda})^2 + ({y\over\lambda})^2} \right) = f {R\over\lambda^3}. $$ {}From this, one sees that the characteristic parameter $fR/\lambda^3$ governs the shape of the bulge as well as its height, and that the width of the bulge is comparable to $\lambda$ when $\lambda{\lower4pt\hbox{$\buildrel<\over\sim$}}\lambda_0$. To summarize: The size and shape of the bulge in the horizon raised by the fluctuation depends on the ratio $\lambda / \lambda_0$. For $\lambda\ll\lambda_0$ the fluctuation raises a bulge much smaller than itself, whereas for $\lambda\gg\lambda_0$ it is (in our Newtonian picture) much larger. In particular, the bulge becomes comparable to the size of the fluctuation precisely when $\lambda \sim \lambda_0$. This conclusion does not depend on the specific profile chosen for the effective mass density of the fluctuation. A delta-function would lead to the same conclusion, as would a dipolar source with vanishing total energy (perhaps more appropriate as a model of a virtual fluctuation of a quantum field). And it appears that full general relativity again yields a similar relationship between scale $\lambda$ and distortion height $h$ if one makes the drastic approximation of spherical symmetry. The formula (20), if taken literally, implies that a fluctuation on scale $\lambda\ll\lambda_0$ induces a distortion of the horizon much greater than its own size. However it seems implausible that such an effect would be present in a fully relativistic setting, where retardation effects would make such extreme ``action at a distance'' by the fluctuation appear very unrealistic, and one would not expect the influence of a fluctuation to extend much beyond its immediate vicinity. If this is correct, then it becomes plausible that the actual perturbations in the horizon due to fluctuations of size $\lambda$ would themselves be of size $\lambda$ for all $\lambda{\lower4pt\hbox{$\buildrel<\over\sim$}}\lambda_0\sim{}M^{1/3}$. The resulting structure of the horizon could then be described as fractal on scales between $1$ and $M^{1/3}$ (it being doubtful whether spacetime itself exists as a continuous manifold on scales below unity). In principle there is no limit to how large this scale-invariant wrinkling could grow if sufficiently massive black holes were available, but unfortunately the prospect of human-sized fluctuations in the horizon disappears when one plugs in the numbers. The wrinkles on a solar mass black hole, for example, would only reach a scale of around $10^{-20}$cm, and for the fluctuations to attain a size of 1cm, a black hole of the absurd mass of $10^{91}$ grams would be called for. \bigskip\noindent In conclusion, I would like to thank R. Salgado for his indispensable aid in preparing the figures. This research was partly supported by NSF grant PHY-9307570. \baselineskip=12pt \vskip 0.5truein \centerline {\bf References} {\partial}\nobreak \medskip \noindent \parindent=0pt \parskip=10pt [1] Schutz, B.F. and R.D.~Sorkin, ``Variational Aspects of Relativistic Field Theories, with Application to Perfect Fluids'', {\it Annals of Phys.} (New York) {\bf 107}:1-43 (1977) [2] J.~Bekenstein, ``Do we understand black hole entropy?'', gr-qc/9409015 [3] R.D.~Sorkin and M. Varadarajan, ``Energy Extremality in the Presence of a Black Hole'', (in preparation) [4] D. Sudarsky and R. Wald, Phys. Rev. D {\bf 46}, 1453 (1990) [5] R.D.~Sorkin, ``Toward an Explanation of Entropy Increase in the Presence of Quantum Black Holes'', {\it Phys. Rev. Lett.} {\bf 56}, 1885-1888 (1986); see also R.D.~Sorkin ``Forks in the Road, on the Way to Quantum Gravity'', talk given at the conference entitled ``Directions in General Relativity'', held at College Park, Maryland, May, 1993, (Syracuse University preprint SU-GP-93-12-2). [6] R.D.~Sorkin, ``Conserved Quantities as Action Variations'', in Isenberg, J.W., (ed.), {\it Mathematics and General Relativity}, pp. 23-37 (Volume 71 in the AMS's Contemporary Mathematics series) (Proceedings of a conference, held June 1986 in Santa Cruz, California) (Providence, American Mathematical Society, 1988) [7] R.D.~Sorkin, ``The Gravitational-Electromagnetic Noether-Operator and the Second-Order Energy Flux'', {\it Proceedings of the Royal Society London A} {\bf 435}, 635-644, (1991) [8] R.D.~Sorkin, ``On Stress-Energy Tensors'', {\it Gen. Rel. Grav.} {\bf 8}:437-449 (1977) [9] Lee, J. and R.D.~Sorkin, ``A Derivation of a Bogomol'ny Inequality in Five-dimensional Kaluza-Klein Theory'', {\it Comm. Math. Phys.} {\bf 116}, 353-364 (1988); Bombelli, L., Koul, R.K., Kunstatter, G., Lee, J. and R.D.~Sorkin, ``On Energy in Five-dimensional Gravity'', {\it Nuc. Phys.} {\bf B289}, 735-756 (1987). [10] R.D.~Sorkin, ``On the Entropy of the Vacuum Outside a Horizon'', in B. Bertotti, F. de Felice, Pascolini, A., (eds.), {\it General Relativity and Gravitation}, vol. II, 734-736 (Roma, Consiglio Nazionale Delle Ricerche, 1983); Bombelli, L., Koul, R.K., Lee, J. and R.D.~Sorkin, ``A Quantum Source of Entropy for Black Holes'', {\it Phys. Rev.} {\bf D34}, 373-383 (1986). [11] See the second reference in [10], p. 374. [12] L.~Susskind and J.~Uglum, {\it Phys. Rev.} D {\bf 50}:2700 (1994) \end
\section{INTRODUCTION} Since the last decade, $\mbox{Fe}_{80-\mbox{x}}\mbox{Ni}_{\mbox{x}}\mbox{Cr}_{20}$ ($14\leq x \leq 30)$ alloys have been the subject of considerable interest because of their diverse magnetic properties within the same crystallographic phase.\cite{akm} These alloys exhibit a compositional phase transition from a long-range AF phase(X=14) to a long-range FM one(X=30), passing through intermediate phases of SG(X=19) and RSG(X=23,26) with the increase of Ni concentration. The presence of the strong competing ferro- [I(Ni-Ni), I(Fe-Ni), I(Ni-Cr), I(Fe-Cr)] and antiferromagnetic[I(Fe-Fe), I(Cr-Cr)] exchange interactions \cite{men's} is responsible for such rich magnetic phases. The complete phase diagram of these alloys was established by Majumdar and Blanckenhagen through dc magnetization and neutron diffraction measurements.\cite{akm} Later on ac susceptibility \cite{sbroy} and magnetoresistance measurements \cite{tkn} also confirmed the proposed phase diagram. Taking advantage of the diverse magnetic properties of these alloys {\em within the same fcc $\mbox{$\gamma$}$-phase}, we tried to understand the age-old problem of relaxation dynamics. This area has always been in the front-line because of its multiple complexities and wide varieties. In this paper, we present the dynamics of time and temperature-dependent magnetic relaxations in various magnetic phases. The search for magnetic relaxation began a century back, when Ewing \cite{ewing} observed persistence of magnetization in soft iron for significant amount of time and a non-exponential decay. Richter \cite{richter} observed logarithmic decay of magnetization for about one decade of time in carbonised iron. Street and Woolley \cite{sw} and Neel\cite{neel} also predicted a logarithmic decay of TRM in FM. Several theoretical and experimental evidences suggest that an anomalous slower relaxation of the form \begin{equation} M(t)=M_0\exp\,[-(t/\tau)^{\beta}]\;,\; \;0<\beta<1\;, \end{equation} is far more reasonable and common than the conventional Debye exponential form($\mbox{$\beta$} =1$). In fact, this kind of relaxation had been observed for a wide range of phenomena and materials.\cite{ngai} In 1970, Williams and Watts \cite{ww} postulated similar functions for dielectric relaxation ($\mbox{$\beta$}$ = 0.5). In a review article Jonscher \cite{jonscher} summarized the experimental evidence on the frequency, time, and temperature dependence of the dielectric response for a wide range of solids. He also found a universality of dielectric behaviour and proposed a generalized approach of many-body interaction. The structural relaxation rate, in the case of liquid to glass transition, can be expressed as a stretched exponential (2/3 $<$ $\mbox{$\beta$}$ $<$ 1). \cite{cohen} Also the validity of this functional form for the relaxation of TRM in SG has been reported by a large number of investigators.\cite{rvc84,rhoo,fmezei,rvc85,nordblad} Palmar et al. \cite{pal} presented in an elegant fashion the whole scenario of similar kinds of relaxation in complex, slowly relaxing, and strongly interacting materials. They considered series relaxation and hierarchically constrained dynamics which is distinctly different from other approaches \cite{mfs} of getting similar results. Hammann et al. proposed a phenomenological picture of the dynamic properties of SG based on fractal cluster model.\cite{ham,ll} Early decay measurements of TRM of spin-glasses had shown logarithmic dependence.\cite{rt} Similar dependence was also observed in $\underline{\mbox{Au}}$Fe, $\underline{\mbox{Ag}}$Mn, $\underline{\mbox{Th}}$Gd spin-glasses from 5 to 10$^4$ s.\cite{so} Analysing the neutron diffraction and ac susceptibility data, Murani \cite{apm} found power law decay for shorter time and logarithmic decay for longer time below T$_{\mbox{g}}$ in SG. Bontemps and Orbach\cite{nbon} measured the TRM of the insulating SG Eu$_{0.4}$Sr$_{0.6}$S between 0.86T$_g$ and 1.04T$_g$. They found power law decay of the TRM for shorter interval of time and a stretched exponential decay beyond a well-defined cross-over time. Recently two different theoretical models have been proposed to describe the SG behaviour. The first is the mean-field approach of Parisis's solution \cite{parisi} of infinite range Sherrington-Kirkpatrick(S-K) \cite{sk} model by considering hierarchical organizations of infinite number of quasi-equilibrium states in phase space.\cite{sibani} The second is the phenomenological approach based on the existence of a distribution of droplets \cite{fisher} or dynamical domains \cite{koper} of correlated spins. Both these theories can explain reasonably well the slower SG dynamics and ageing effects. A nice comparison between these two models was given by Lefloch et al.\cite{lef} Huse and Fisher \cite{fisher} proposed long time decay as stretched exponential with exponent 1/2 by using droplet fluctuations in two dimensional pure Ising system with a spontaneously broken continuous symmetry. In the framework of droplet fluctuations,\cite{fisher} the spin autocorrelation can be described as exponentially rare in $\ln t$ : $\overline{\mbox{c}_{\mbox{i}}(\mbox{t})}$ = $\exp[-k(\ln t)^{y}]$ (where $y = 1$ for random exchanges, $y = d-2/d-1$ for random fields) in random FM and power of $\ln t$ in SG. Ogielski\cite{ogi} predicted that the spin autocorrelation function can be described by the product of power law and stretched exponential at all temperatures above T$_g$\cite{gunn} and by power law below T$_g$. Ocio et al. found an analytical expression based on scaling analysis of ageing effects, for the decay of TRM in SG, CsNiFeF$_6$ \cite{ocio} as well as in AgMn.\cite{alba} They proposed the decay of TRM as \begin{equation} M(t)=M_0\;(\lambda)^{-\alpha}\,\exp\,[-(\omega\, t/t_{w}^{\mu})^ {1-n}]\;,\; \end{equation} \begin{equation} with\,\,\;\lambda=(t_{w}/(1-\mu))[(1+t/t_{w})^{1-\mu}-1] \;, \end{equation} where $\mbox{$\lambda$}$ is the effective time and $\mbox{$\mu$}$ is an exponent smaller than 1 . A different analysis, based on S-K mean-field model, \cite{hs} suggests algebraic decay. In ferromagnets attempts have been made to explain the magnetic relaxation using a model based on magnon relaxation on a percolation distribution of finite domains. \cite{rvc90} The other popular prediction of relaxation is the power law \cite{rjb} decay which can be obtained from scaling theories for domain growth \cite{gunton} and internal dynamics.\cite{halp} Ikeda and Kikuta\cite{ike} found that there is no magnetic relaxation in AF over 10 h in $Mn_{0.45}Zn_{0.55}F_2$ at a slightly lower temperature than the transition temperature. Despite considerable experimental work on relaxation dynamics covering enormous range of time window, no conclusive results have been found. Moreover, there is a lack of clear distinction between the SG and the RSG phases. Also no experimental data are available on relaxation dynamics in the AF phase. All these have motivated us to study systematically the relaxation dynamics in four different magnetic phases, namely, SG, RSG, FM and AF in FeNiCr alloys {\em within the same crystallographic phase}, for different wait times and at different temperatures for the largest available time window. There is a running controversy about the existence of the FM orderings in RSG at the lowest temperature as predicted by the GT model.\cite{gt} We have also measured linear and non-linear ac susceptibilities for the sample X=23 to resolve this controversy. \section{EXPERIMENTAL PROCEDURE} All these ternary alloys were prepared by induction melting in argon atmosphere. The starting materials were of 99.999\% purity obtained from M/s Johnson Mathey Inc., England. The alloys were cut to the required size, homogenized at 1323 K for 100 h in argon atmosphere, and then quenched in oil. Chemical analysis of Ni and Cr shows that the compositions of the alloys are within $\pm 0.5$ at. $\%$ of their nominal values. X-ray diffraction data at room temperature in powdered samples reveal that these are single-phase fcc ($\gamma $) alloys with the lattice parameter a=3.60 $A^{o}$. Neutron diffraction data show the presence of single-phase fcc ($\gamma $)structure down to 2 K for X=19 alloy. The two most essential ingredients of SG are "frustration" and "quenched disorder" which manifest themselves in history dependent phenomena and eventually lead to nonequilibrium behaviour below T$_{\mbox{g}}$.\cite{binder} To study the relaxation dynamics, we applied a small magnetic field (10 gauss) at a temperature greater than the characteristic temperatures (T$_{\mbox{g}}$ , T$_{\mbox{c}}$ , T$_N$ ) , cool ($\mbox{t}_{\mbox{c}}$ = cooling time ) the system to a temperature T$_{\mbox{m}}$ which is less than the characteristic temperature. This field-cooled system will attain a metastable state (which is not an equilibrium state), \cite{rhoo,rvcprb84} then after waiting for definite amounts of time $\mbox{t}_{\mbox{w}}$ ( $\mbox{t}_{\mbox{w}}$ varies from 60 to 3600 s ) the magnetic field is removed, the system allowed to relax towards equilibrium state and then TRM measured with time (till $\sim 10^{4}$ s) using a SQUID magnetometer (Quantum Design MPMS). To see the effect of temperature, we also measured TRM at several temperatures, T$_{\mbox{m}}$, both above and below the characteristic temperatures for constant wait time, t$_{\mbox{w}}$ = 180 s. We have very carefully subtracted the background noise. We repeated the experiments and obtained similar results within the experimental error. \section{RESULTS AND DISCUSSION} We observe remarkable results of ageing effects in the SG where each isotherm strongly depends upon wait time t$_{\mbox{w}}$ (time of exposure in magnetic field below T$_{\mbox{g}}$). The magnetization can be well represented by an equation of the form \begin{equation} M(t)=M_0(t/t_{w})^{-\gamma}\exp\,[-(t/\tau)^{1-n}] \end{equation} for the entire time domain. This is a simpler version of the earlier prediction of scaling analysis of ageing process \cite{alba,ocio} (Eq. (2)). Instead of t they used $\lambda$ which is a function t and t$_w$ and $\lambda$ $\rightarrow$ t for t $<<$ t$_w$. But we find that the Eq. (4) is valid for the entire time domain. Ogielski\cite{ogi} predicted similar analytical form to describe the spin autocorrelation function in SG at all temperatures above T$_g$. However other investigators \cite{rvc84,rvc85,nordblad,chu} found only the stretched exponential form for the decay of TRM for the SG phase. We observe in the RSG the best representation of TRM is \begin{equation} M(t)=M_{1}+ M_0\,\exp\,[-(t/\tau)^{1-n}]\,\;. \end{equation} The additional small term $\mbox{M}_1$ can easily be explained in the framework of the GT model \cite{gt} where only transverse spin freezing occurs in the RSG while the longitudinal spins can produce diffuse background effect. Similar expression also reported for the decay of TRM in SG as well as in RSG.\cite{p-m} They did not find any difference between the RSG and the SG phases. We find that Eq. (5) is only valid for the RSG and distinct differences exist between the SG and the RSG phases. The most salient feature of the present work is that it can demonstrate how changing the composition by small amounts in $\mbox{Fe}_{80-\mbox{x}}\mbox{Ni}_{\mbox{x}}\mbox{Cr}_{20}$ alloys ( X = 14 AF, X = 19 SG, X = 23,26 RSG, X = 30 FM ) one can tune the relaxation dynamics. This is the only experimental report which presents a complete scenario of the relaxation spectrum in various kinds of interesting magnetic phases and throws new light on this area of interest to theoreticians as well as experimentalists . Figure 1 shows the time decay of TRM, M(t), for different wait times (t$_{\mbox{w}}$ = 60, 240, 1200, 1800, 3600) below T$_{\mbox{g}}$(12 K) (T$_{\mbox{m}}$ = 5 K) for the SG (X = 19) and the solid lines are the best fits of the experimental data to Eq.(4) for almost four decades of time (till 12,000 s) ($\chi^2=(1/n)$\(\sum_{i=1}^{n}(Raw\;data_i-Fitted\;data_i)^2/Raw\;data_i^2\)$ \leq 10^{-6}$). We have purposefully plotted our data on linear time scale because if we plot on log scale, for longer time interval, the plot will contract and the fits will apprently look better. However, goodness of fit is better judged from the value of $\chi^2$. From these fits, the value of the initial TRM, $\mbox{M}_0$ , the characteristic time constant, $\mbox{$\tau$}$, and the exponents, n and $\mbox{$\gamma$}$, are found. The most significant feature of this analysis is that M$_0$ ($\approx$ 0.04 emu/g) is not varying with t$_{\mbox{w}}$ \cite{rvc84,rhoo,alba} while the exponent, n, gradually increases with the increase of t$_{\mbox{w}}$. This indicates that larger time exposure in magnetic field below T$_{\mbox{g}}$ makes the system more reluctant to come back to an equilibrium state from a metastable one. Other investigators \cite{rvc84,rhoo} reported n independent of wait time but we found that it varies from 0.63 to 0.77. The power law exponent, $\mbox{$\gamma$}$, remains constant (0.022 $\pm$ 0.004) except $\mbox{t}_{\mbox{w}}$ = 60 s where $\mbox{$\gamma$}$ is 0.03. Figure 2 shows the variation of TRM , M (t), with time for different temperatures at constant wait time ($\mbox{t}_{\mbox{w}}$ =180 s). From the best fit to Eq.(4) the temperature variations of n, $\mbox{M}_0$, $\mbox{$\tau$}$, and $\mbox{$\gamma$}$ are found which broadly match with the previous observations.\cite{rvc84,rhoo,alba,chu} The value of n increases linearly from 0.8 to 0.9 from 0.5T$_{\mbox{g}}$ to T$_{\mbox{g}}$ and then it starts falling beyond T$_{\mbox{g}}$ (Fig. 3). It is reported that n remains constant at lower temperatures and then starts rising from a temperature T =T$_0$, which strongly depends on the anisotropy energy of the sample \cite{rvc85} and hence can vary from sample to sample. Non-availability of lower temperature data(less than 0.4T$_{\mbox{g}}$) prevented us from verifying the constancy of n. The prefactor, M$_0$, shows (Fig. 3) a linear decrease with increasing temperature for T $<$ 0.75 T$_{\mbox{g}}$ and the rate of decrease becomes slower for T $>$ 0.75 T$_{\mbox{g}}$, which is in good agreement with the earlier findings.\cite{rvc84} The power law exponent $\mbox{$\gamma$}$ increases with the increase of temperature up to 0.75 T$_{\mbox{g}}$. At T$_{\mbox{g}}$ and beyond it starts decreasing with temperature (Fig.~3). It is difficult to tell exactly the temperature from which it has started falling , but earlier prediction was that it should increase as one approaches T$_{\mbox{g}}$.\cite{alba} Figure 4 shows how the inverse of the characteristic relaxation time $\mbox{$\tau$}$ decreases with the increase of the reduced temperature T$_{\mbox{g}}$/T, as observed by Alba et al.\cite{alba} It also reflects the general tendency observed by others.\cite{rhoo,chu} Fewer number of data points prevents us from finding the functional form. Apparently, to check the exponential form as predicted earlier\cite{rhoo} we need to probe even at lower temperatures. We can try to explain the variation of $\mbox{$\tau$}$ with temperature by considering hierarchically organized metastable states in phase space \cite{lef} within the framework of Parisi's mean-field solution of infinite range S-K model. The states are continuously splitting into new states with the lowering of temperature and are separated by barriers. These barrier heights strongly vary inversely with temperature. That is, at lower temperatures, barrier heights increase and separate different metastable states into mutually inaccessible states which makes $\mbox{$\tau$}$ larger at lower temperatures. Using values of n and $\mbox{$\tau$}$ from Figs. 3 and 4, respectively and plotting $\log(1-n)(1/\tau)^{1-n}$ versus (1-n) by writing this function as $(1-n)(1/\tau)^{1-n}$ = $c\, \omega^{1-n}$, we get the value of the relaxation frequency $\mbox{$\omega$}$ from the slope of the best fitted curve (Fig.5) as 1.9$\times10^{-7}$ s$^{-1}$ and the constant, c, is 0.13. These values are much less than those predicted earlier.\cite{rvc84} Analysis of M(t) by considering only the stretched exponential, ie., without the power law part of Eq.(4), shows unusually small ($\approx 10 $ times smaller) values of n and a poor $\mbox{$\chi$}^2$($\approx 10^{-5}$). Fits are even poorer in cases of other mathematical forms, like the power law. So, we find that Eq.(4) is the simplest analytical form that can represents our experimental data of the decay of TRM in the SG phase for the entire time domain. The SG has been the focus of attention for quite sometime, but not much attention has been paid to the RSG phase. The sample with X = 23, below 35 K, enters a FM phase from a random paramagnetic(PM) phase. On further lowering of temperature below 22 K it enters once again a new random phase where FM and SG orderings coexist.\cite{akm} It shows most of the SG-like behaviour (frustration, irreversibility, etc. ). This phase is known as the RSG. Figure 6 represents the variation of M(t) with time at T$_m$=5 K for different wait times (t$_{\mbox{w}}$ = 60, 1200, 1800, and 3600 s ) in the RSG phase (X=23) and the solid lines are the best fitted curves which are of the form of Eq.(5)($\chi^2 \approx 10^{-6}-10^{-7}$). The salient feature of this figure is that the initial magnetization, $\mbox{$\sigma$}_0$ = M$_0$ + M$_1$, increases with the increase of wait time (Fig. 7). This is not observed in the SG phase but the variation of n is similar. We also observe that M$_1$, which is arising because of the presence of the FM ordering, \cite{gt} is not changing at all with wait time and the values of M$_1$ are 84$\%$ to 94$\%$ of the total magnetization depending on wait time (total magnetization changing with wait time). It is quite natural that if the ferromagnetic component is embedded in the SG phase, then the major contribution of the total magnetization should come from the ferromagnetic component. Figure 8 depicts the variation of M(t) with time at different temperatures ( T$_{\mbox{m}}$ = 10, 15, 20, 25, 30, 38 K ) for constant wait time (180 s). From the fits of the data to Eq. (5), the value of $\mbox{$\sigma$}_0$ and the exponent, n, are found. The exponent increases with temperature and approaches unity as the temperature approaches T$_{\mbox{g}}$. This is exactly in agreement with the previous observations \cite{rvc84,rhoo} in SG, the only difference is that the values are slightly higher. If we increase the temperature even more, a sudden change takes place around 20~K. A drop in the value of n indicates a phase change. Moreover, at lower temperatures it shows better fit to Eq.~(5) but at higher temperatures power law fit is better than the stretched exponential. This supports the on-set of the FM phase. This could also be an indication of the switch-over from nonequilibrium dynamics to equilibrium dynamics when it passes from the RSG to the FM phase. The power law behaviour in the FM phase (having an exponent $\approx$ 0.06) is quite consistent with the Huse and Fisher theory.\cite{fisher} The rate of increase of the exponent also changes somewhat beyond 30 K where it passes from the FM to the PM state at 35 K. The value of the exponent varies by 4.6 $\%$ in the temperature interval of 8 K (30 K to 38 K)(Fig. 9). This is above the error bar which is less than 1 $\%$. The size of the symbol (Fig. 9) is of the order of the error bar. The value of the exponent shows anomaly near two temperatures, 22 K and 35 K, which are nothing but T$_{\mbox{g}}$ and T$_{\mbox{c}}$, respectively.\cite{akm} But the variation of n near T$_c$ is not as prominent as that of near T$_g$. We also find that in Eq.~(5) the additional term , M$_1$, which is the value of the residual magnetization, M($\infty$), decreases with the increase of temperature below T$_{\mbox{g}}$ (0.0067 emu/g at T$_{\mbox{g}}$) and suddenly increases to 0.027 emu/g when it enters the FM phase, as expected. The most interesting observation is the variation of the initial magnetization, $\sigma_0$, with temperature (Fig. 9). $\mbox{$\sigma$}_0$ decreases monotonically with the increase of temperature up to 20 K beyond which the rate of decrease reduces significantly and at 25 K and 30 K it becomes almost constant (0.033 and 0.032 emu/g, respectively) and then there is a sudden rise at 38 K (0.045 emu/g) which makes the scenario most interesting. The value of $\sigma_0$ changes about 40$\%$ in this temperature interval (30 to 38 K). This kind of remarkable observation of switching of magnetization while passing from FM to PM phase was reported earlier only by Chamberlin and Holtzberg \cite{rvc91} in ferromagnetic EuS single crystal. They tried to explain this in terms of the percolation theory.\cite{stauffer} So we observe that the TRM in RSG (X=23) shows a minimum near T$_c$ and a local maximum just beyond T$_{\mbox{c}}$. At lower temperatures ($<$ T$_{\mbox{c}}$), the finite size domains try to orient themselves with the direction of the local field, which need not be in the same direction as the applied field. These domains are dynamically strongly correlated (forming a strong viscous medium). FeNiCr alloys have shown very large high-field susceptibility,\cite{mit} i.e., even at very high magnetic fields the orientation of these domains with the direction of the applied field is not complete because of the presence of strong anisotropy. Near T$_{\mbox{c}}$ the correlation between finite domains gets disrupted. Just above T$_{\mbox{c}}$, the domain magnetization still remains but the domains become less viscous and they try to orient themselves along the direction of the applied field, thus increasing the magnetization. Further increase of temperature randomizes the whole spin orientation. We repeated the experiment to confirm this unusual observation and found similar behaviour. In general the non-linear magnetization in the presence of a magnetic field(h), can be written as $ m = m_{0}+\chi_{0}h + \chi_{1}h^2 + \chi_{2}h^3 + ....$, where $m_0$=spontaneous magnetisation, $\chi_0$ linear- and $\chi_{1},\chi_{2}, etc.,$ are nonlinear susceptibilities, and $h=h_0\; Sin\; \omega t$ for a.c field. In SG where no spontaneous magnetization exists, only odd harmonics of the susceptibility will be present. \cite{tosi,masuo} We observe that the out-of-phase part of the linear a.c susceptibility ($\chi_{0}^{\prime\prime}$) shows peaks \cite{masuo} near T$_{\mbox{g}}$ and T$_{\mbox{c}}$, respectively (Fig. 10) (a.c field of 0.6 Oe and frequency $\omega$ =242 Hz). Similar peaks are observed in $\chi_{2}^{\prime\prime}$ ($\partial^3m/\partial h^3$, out-of-phase part of the 3rd harmonics). The distinct peak near T$_{\mbox{g}}$ for X=23 is consistent with the theoretical predictions.\cite{masuo,e-a} $\chi_{1}$ ($\partial^2m/\partial h^2$, 2nd harmonics) also shows a distinct peak near T$_{\mbox{g}}$ (Fig. 10) which is never observed in pure SG, including X=19. This indicate the presence of a ferromagnetic component in RSG down to the lowest temperature. So, we find that the RSG shows the SG transition and also the presence of FM ordering below T$_{\mbox{g}}$. This distinguishes it from the SG. Detailed studies of ac susceptibilities in FeNiCr alloys will be published elsewhere. The sample with X=26, below 56 K , enters a FM phase from a random PM phase. On further lowering of temperature below 7 K it enters a RSG phase.\cite{akm} We observe some different features in the samples X=23 and X=26, though they undergoes similar kinds of phase transitions(PM $\rightarrow $ FM $\rightarrow$ RSG). M(t) data at 5 K, similar to those shown in Fig. (6) but now for the sample X=26, when fitted to Eq. (5) ($\chi^2\approx 10^{-6}-10 ^{-7}$), shows that the initial magnetization, $\mbox{$\sigma$}_0$ = M$_0$ + M$_1$, does not change with wait time ($\approx 0.5 $ emu/g). The value of the exponent, n, increases from 0.55 to 0.64, with the increase of wait time from 240 s to 3780 s. We have observed similar variation in the SG(X=19) phase whereas in the other sample, X=23, $\mbox{$\sigma$}_0$ increases with wait time in the RSG phase. The value of $\mbox{$\sigma$}_0$ is larger for the sample X=26 than that for the sample X=23 and smaller than that for the sample X=30(FM, describe below). Hence the value of the initial magnetization, $\mbox{$\sigma$}_0$, increases gradually with the increase of Ni concentration as we move towards the FM phase. This is exactly in agreement with the previous observations.\cite{akm} Variation of n is similar to both the SG(X=19) and the RSG(X=23) samples. Figure 11 displays the variation of M(t) with time at different temperatures (T$_{\mbox{m}}$ = 6, 8, 10, 20, 30, 40, 50, and 60 K) for constant wait time (180 s)for the sample X=26. From the fits ($\chi^2\approx 10^{-6}-10 ^{-7}$) of the data to Eq. (5), the values of initial magnetization, $\mbox{$\sigma$}_0$, and the exponent, n, are found. $\mbox{$\sigma$}_0$ decreases at a faster rate up to 20 K and then continues to decrease slowly till 60 K (Fig. 12). We have not observed the local maxima above T$_{\mbox{c}}$ as in the case of X=23. To observe this we need to probe at a temperature closer to T$_{\mbox{c}}$(56 K). The exponent, n, increases abruptly when the system undergoes a phase transition from the RSG to the FM at 7 K. Then it starts to decrease up to 20 K and beyond this again increases till 50 K. Further increase of temperature reduce the value of n and the system passes from the FM to the PM phase (Fig. 12). The variation of n is not well understood, specially the dip around 20 K. Interestingly, we have also observed around this temperature some striking features in the low-field magnetoresistance and a.c susceptibility measurements.\cite{tbp} These have kept the field wide open for further work. We also observe that at higher temperature (T$\geq$30 K, much higher than T$_g$) M(t) data, show better fits to the power law compared to the stretched exponential (Eq. (5)). For X=23, we observe similar behaviour for T $\geq$ T$_g$. Figure 13 shows the variation of TRM with time at 5 K in the FM phase (X=30) for different wait times (t$_{\mbox{w}}$ = 240, 1380, 1980, 3780 s). We find that \begin{equation} M(t) = M_1 + M_0\;t^{-\gamma} \end{equation} is the best fit of the experimental data (solid lines in the graph) and from the fits ($\chi^2\approx 10^{-6}-10 ^{-7}$) the values of $M_1$, $M_0$ and $\gamma$ are found. It does not show much wait-time dependence, the values of $M_1$ and $M_0$ remain almost constant ($\approx$ 0.29 and 0.35 $\pm$ 0.008 emu/g, respectively) while the exponent, $\gamma$, decreases inversely with wait time (from 0.049 to 0.036). With the increase of temperature the value of the initial magnetization reduces drastically (0.65 emu/g at 5 K and 0.094 emu/g at 80 K) and the exponent also becomes smaller (0.049 at 5 K and 0.00036 at 80 K). Figure 14 shows the time decay of TRM, M(t), for different wait times (t$_{\mbox{w}}$ = 180, 1320, 1980, 3780 s) below T$_N$=26~K (T$_{\mbox{m}}$ = 5 K) for the AF(X = 14) sample and the solid lines are the power law fits of the form \begin{equation} M(t) = M_0 t^{-\gamma}. \end{equation} From these fits the value of $M_0$ and $\gamma$ are obtained (Table I). $M_0$ ($\approx$ 0.00206 emu/g) does not change with wait time whereas $\gamma$ decreases with the increase of wait time (0.0044 to 0.0038). Figure 15 shows the time decay of TRM, M(t), at different temperatures for constant wait time (180 s) and the solid lines are the power law fits from which the value of $M_0$ and $\gamma$ are found (Table III). The value of $M_0$ decreases linearly with the increase of temperature up to T$_N$ (0.0015 emu/g at 24 K) and then suddenly falls to a much lower value (0.00012 emu/g at 30 K) as shown in Fig. 16. The exponent, $\gamma$, increases with temperature and the rate of increase changes somewhat when it passes to the PM phase (Fig. 16). In case of the AF phase we need not add any constant term (unlike the FM phase) and the value of the exponent is an order of magnitude lower than that in the FM phase. We also observe that the stretched exponential function, Eq.(1), shows reasonably good fits to the TRM in the AF phase. The value of the fitted parameters and the $\chi^2$ are given in Table II and IV. The value of the exponent, $\mbox{$\beta$}$, is almost two orders of magnitude smaller than that in the SG phase (0.004 and 0.37, respectively). It increases monotonically (Table IV) with temperature (0.004 at 5 K and 0.029 at 30 K) in contrast to that in the SG phase where $\mbox{$\beta$}$ decreases with the increase of temperature up to T$_ {\mbox{g}}$ (0.37 at 5 K and 0.1 at 12 K obtained from Fig. 3 using $\beta$ = 1-n). So, we find that the M(t) data for AF fit well to both the power law (Eq. (7)) and the stretched exponential (Eq. (1)). We have given the values of $\chi^2$ in Table I-IV for comparison. They are comparable for both the above mathematical forms. Hence it is difficult to describe the exact nature of the decay of the TRM in the AF phase. More experimental work is needed to arrive at a more definite conclusion. \section{conclusion} We have measured the TRM in $\mbox{Fe}_{80-\mbox{x}}\mbox{Ni}_{\mbox{x}}\mbox{Cr}_{20}$ ($14\leq \mbox{x} \leq 30$) alloys for four different magnetic phases within the {\em {same crystallographic phase}} and from their wait time and temperature variations tried to establish a correspondence with the phase diagram. We find distinct differences between the SG and the RSG phases with two different analytical forms for the time decay of the TRM. We also observe the presence of the FM ordering in the RSG below T$_{\mbox{g}}$ which is consistent with the GT model. The peak in $\chi_1$ near T$_{\mbox{g}}$ for RSG (X=23) had never been observed in any pure SG, including X=19. The peak can only be observed if spontaneous magnetization is present. Hence we confirm the presence of the FM ordering in RSG down to the lowest temperature. This feature distinguishes the RSG from the SG. We also report the remarkable observation of the local maximum of the TRM just above T$_{\mbox{c}}$ in the RSG (X=23) when it passes to the PM phase from the FM phase. This is found for the first time in any polycrystalline RSG. However, it's exact theoretical justification is unclear. We also observe that the value of the exponents show anomaly near the phase transitions. We observe some different features in the samples X=23 and X=26, though they undergoes similar kinds of phase transitions. We find the conventional power law decay in the FM phase. The value of the magnetization is found to increase with the Ni concentration, as predicted earlier. In the AF phase the power law decay is indistinguishable from the stretched exponential as a description of the TRM. More experimental work is needed to arrive at a more definite conclusion about the decay of TRM in AF. \section{Acknowledgement} Financial assistance from project No. SP/S2/M-45/89 of the Department of Science and Technology, Government of India, is gratefully acknowledged.
\section{Introduction} \setcounter{equation}{0} This paper deals with the motion of a particle in a correlated random potential under the influence of a driving force. The correlations of the disorder are characterized by a short distance cutoff and a power law decay at larger distances with exponent $\gamma$. This problem has been solved in one dimension \cite{Sinai82,leDou95,Scheidl95} and, depending on $\gamma$, drift with finite mobility, creep or pinning was found. Creep means that the particle moves with a mean velocity less than proportional to the driving force and in case of pinning the mean velocity is zero, unless the force exceeds some critical value. In $N$ dimensions a mean field treatment becomes exact for $N\to\infty$ and there is a close formal similarity to spin glass problems with long range interactions. The situation without external driving force has been treated within replica theory \cite{Mez:manifold90,Mez:manifold91} and also using a stochastic dynamics approach \cite{KibachHo93I,KibachHo93II}. Actually the more general case of the motion of a $D$-dimensional manifold in an $N\to\infty$ dimensional space was investigated. For finite $N$ the mean field treatment is only an approximation. Nevertheless it might give some clue about a great variety of systems with disorder, for instance flux lines in superconductors or interfaces in random field systems. The close formal relation to spin glass problems raises another set of questions. There has been a continuous interest in the dynamics of spin glasses and related systems. The pioneering work of Sompolinsky and Zippelius \cite{Somp81,SZ82} concentrated in part on an understanding of the replica theory and the hierarchical replica symmetry breaking scheme proposed by Parisi \cite{Par79}. The basic assumption was about the existence of a hierarchy of diverging time scales with ultrametric properties. This point has been investigated further \cite{Ho:DySK84} introducing a long time scale $\bar t$ associated with slow changes of the random interactions in this model. It could be shown that the above assumption meant that the long time contributions to the correlation and response functions have to be smooth functions of $x(t)=1-\ln t/\ln\bar t$ with $\bar t\to\infty$. It was later shown \cite{Ho:DySK87} that this led to inconsistencies which could not be resolved at that time. There exists a variety of models with disorder where a replica treatment requires only single step replica symmetry breaking instead of the hierarchical scheme mentioned above. The resulting phase transition can be considered as discontinuous in contrast to the continuous transition observed in the SK-model and other cases requiring hierarchical replica symmetry breaking. A treatment via dynamics \cite{KibachHo93I,KiTh87,ho:binperc92,CHS93} revealed a dynamic freezing temperature above the transition temperature obtained in replica theory. This is in contrast to the continuous transitions where the same transition temperature is found. Furthermore it could be shown \cite{CHS93} that the states which contribute to dynamics have a higher energy than the states relevant for the replica calculation. Numerical simulations of a learning process in a perceptron with binary bonds \cite{ho:binperc92}, another system with discontinuous transition, indicated that the replica result applies if the limit $\bar t\to\infty$ is performed first and then the limit $N\to\infty$, whereas the results obtained via dynamics hold for the opposite order. This observation suggests that in the thermodynamic limit the particular states most relevant for the replica calculation cannot be reached within any finite time. This picture is consistent with the fact that finding the best groundstate in a spin glass or perfect learning in a perceptron with binary bonds is a combinatorial hard problem which cannot be solved in polynomial time. The replica treatment, on the other hand, is purely static and does not refer to any kind of dynamics. For a given physical situation the appropriate order of limits has of course to be found, keeping in mind that the longest equilibration times are likely to diverge exponentially with some power of $N$ \cite{KibachHo91}. This raises the question whether similar discrepancies also show up in systems with continuous transition. The problem of the motion of a particle in a random potential under the influence of a driving force is in several aspects an appropriate model system to address this question. Prescribing the drift velocity $v$ and calculating the force necessary to drive the particle, a long external time scale is defined by $\bar t\sim v^{-1}$. For times exceeding $\bar t$ the system is supposed to reach a stationary non-equilibrium state, a so-called flow state, where correlation and response functions depend on differences of time only. This is of great advantage regarding numerical as well as analytic work. Depending on the exponent $\gamma$ and on temperature, a continuous as well as a discontinuous transition can be found and therefore both cases can be studied on a single model. The results presented in the following are based on numerical integrations of the dynamic mean field equations of this model, extending over more than thirty orders of magnitude in time and velocity. Such a wide range is actually necessary in order to understand the different scaling and crossover regimes emerging in this problem. Such a wide range of time scales is not unrealistic keeping in mind that phenomena might be observed on a scale of days or more with an intrinsic time scale of $10^{-12}$ to $10^{-16}$ seconds typical for vibrational or electronic degrees of freedom. The numerical results are supplemented by exhaustive analytic work characterizing the different scaling regimes and focusing on their asymptotic properties and associated exponents. The characteristic time scales are obtained by matching between adjacent scaling regimes and are again governed by characteristic exponents. The key result of the present investigation is the observation that the dynamics with a long but finite external time scale $\bar t$ is ruled by three regimes: \begin{description \item[i)] The FDT-regime describing local equilibrium at short times. \item[ii)] The intermediate plateau regime for times $t\sim {\bar t}^{\,\xi}$ with $0<\xi<1$. \item[iii)] The asymptotic regime for $t\sim \bar t$. \end{description} These regimes have additional substructures and associated additional internal time scales. The plateau regime is the crucial link between short times and long times of the order of the external time scale. Its properties determine for instance whether creep or pinning is found or whether the longest internal time scale is proportional to $\bar t$ or $\bar t^{\,1+\eta}$ with $\eta>0$. This question is of relevance for instance in the context of aging phenomena \cite{aging1,aging2,aging3}. \mfigure{Phase-D} {Phase diagram. $\cal O$: Drift phase with $v\sim F$ (driving force $F$, resulting velocity $v$). $\cal A$: Creep phase with $v\sim F^{1/\eta}$. $\cal B$ and $\cal C$: Pinning phases with $v=0$ for $F<F_p$. The phase diagram resulting from replica theory is also shown (doted lines)}{phase-fig} Depending on $\gamma$ and temperature, different phases are found. The phase diagram resulting from the present investigations is shown in Fig.\ref{phase-fig}. There is a phase with finite mobility ($\cal O$), another one with creep behavior ($\cal A$) and two more where pinning is observed ($\cal B$ and $\cal C$). This is in qualitative agreement with the findings for the one-dimensional model \cite{leDou95}, which indicates that the transitions are not an artifact of the mean field treatment. In all phases, except the high temperature phase $\cal O$, additional intermediate time scales exist, which diverge for $v\to0$. There is, however, no ultrametric organization of time scales. This is in contrast to the common proposal \cite{KibachHo93II,SZ82,Ho:DySK84,CHS93}. Moreover, the phase diagram suggested by replica theory \cite{Mez:manifold90,Mez:manifold91} or dynamics within the hypothesis of ultrametric time scales \cite{KibachHo93I,KibachHo93II} differs from the one obtained in the present investigation. For instance, $\gamma=1$ has been predicted for the $\cal A$-$\cal B$ phase boundary in contrast to $\gamma=1.3044$ found at present. This leads to the conclusion that the states dominating the replica theory with broken symmetry are again not the same as those relevant for dynamics on long but finite time scales. As a consequence of the assumption of ultrametricity in the organization of diverging time scales, the resulting correlation and response functions were not unique \cite{SZ82,Ho:DySK84}. The present investigation yields on the other hand unique results and this problem is therefore removed. The inconsistencies \cite{Ho:DySK87} mentioned above are also resolved in the sense that the proposed reparametrization of time $x(t)=1-\ln t/\ln\bar t$ is not appropriate. There has recently been a growing interest in the dynamics of glassy systems evolving from a non-equilibrium initial state \cite{CuKu93,FrMe94a,FrMe94b,CuKu94,BCuKuPa95,CuDe95,CuDou95}. The situation there is more complicated because the problem is no longer homogeneous in time and response and correlation functions depend on two time arguments $t$ and $t'$. If, however, $t-t'\ll t'$, one expects a behavior similar to the one investigated at present with $t'$ taking over the role of $\bar t$. Especially if both $t-t'$ and $t'$ diverge with $(t-t')/t'\to0$, one should also expect the appearance of intermediate time scales and associated scaling regimes. This possibility has not been considered at full depth in the above work. For short time $t-t'$ correlation and response functions are related by fluctuation-dissipation theorems (FDT). This is also the case in the present investigation. The different phases have, however, different scenarios of how FDT's are violated at longer time scales. Since this happens for times $1\ll t\ll \bar t$, similar phases are expected for the above mentioned relaxation processes. In this context the non-equilibrium dynamics of the spherical SK-model investigated recently by Cugliandolo and Dean \cite {CuDe95} is of great interest, because this model does not require replica symmetry breaking in its low temperature phase. The present paper is organized as follows: Section 2 contains the definition of the model and the dynamic mean field equations, which have the form of coupled nonlinear integro-differential equations. It also contains the definition of effective time dependent exponents, which are widely used in the following. Section 3 is devoted to the numerical integration of the dynamic mean field equations, to the presentation of results and to a preliminary identification of different time scales and scaling regimes. Section 4 contains analytic results and estimates regarding time scales and scaling properties. It starts with the FDT-solution valid for short time and continues with a discussion of various convolution type integrals entering the self-energies. In bypassing the QFDT-solution \cite{KibachHo93I} and the hierarchical solution \cite{KibachHo93II} are brought up. It is then shown how to evaluate the above integrals using the time dependent effective exponents. This leads to a reformulation of the original mean field equations in terms of a set of 15 coupled ordinary and first order differential equations. This appears complicated but it allows a discussion of different scaling regimes in which only closed subsets of these equations are relevant. These regimes are in the order of increasing time the FDT-regime, the plateau regime, where the correlation function $q(t)$ stays close to the asymptotic value $q_c$ of the FDT-solution, and finally the asymptotic regime for times of the order of the external time scale $v^{-1}$. A concluding discussion follows in Section 5 and a brief derivation of the dynamic mean field equations is given in an appendix. \section{Formulation of the problem and dynamic mean field theory} \setcounter{equation}{0} \subsection{The model} A particle is investigated which moves under the influence of an external force $\bm{F}=\big\{\sqrt{N}F,0,\cdots,0\big\}$ in a random potential $V(\bm{\varrho})$. The coordinates of the particle are written as ${\bm{\varrho}=\big\{\varrho_1+\sqrt{N}vt,\varrho_2,\cdots,\varrho_N\big\}}$ assuming a mean velocity $\bm{v}=\big\{\sqrt{N}v,0,\cdots,0\big\}$. This means that the components $\varrho_i$ are measured in a frame moving with velocity $\sqrt{N}v$ along the $1$-direction. The scaling with $\sqrt{N}$ ensures a nontrivial limit $N\to\infty$. The system is at a temperature $T=\beta^{-1}$ and its Hamiltonian is \begin{equation} H=V(\bm{\varrho})+\sfrac12\mu_0\bm{\varrho}^2-\bm{F}\!\!\cdot\!\!\bm{\varrho} \,,\end{equation} where a confining potential $\sfrac12\mu_0\bm{\varrho}^2$ might be added to the random potential and the potential of the driving force. The following investigation is, however, restricted primarily to $\mu_0=0$. The motion of the particle is governed by a Langevin equation \begin{equation} \partial_t\varrho_{\alpha}=-\beta\frac{\delta H}{\delta \varrho_{\alpha}}+\xi_{\alpha} \end{equation} with Gaussian white noise \begin{eqnarray} \av{\xi_{\alpha}(t)\xi_{\beta}(t')}=2\,\delta_{\alpha\beta}\delta(t-t')\,. \end{eqnarray} The quenched disorder is also assumed to be Gaussian with \begin{eqnarray} &&\overline{V(\bm{\varrho})}=0\nonumber\\ &&\overline{V(\bm{\varrho})V(\bm{\varrho'})}=-N\,f\big({\textstyle\frac 1 N} (\bm{\varrho}-\bm{\varrho'})^2\,\big) \label{VV-av}\end{eqnarray} and \begin{equation} f(x)=\frac 1 {2(1-\gamma)}\,(1+x)^{1-\gamma} \label{f-def}\,.\end{equation} The strength and the short range cutoff of the disorder as well as the diffusion constant are set to one. This model differs from the one studied earlier \cite{Mez:manifold90,Mez:manifold91,KibachHo93I,KibachHo93II} only by the addition of the driving force. \subsection{Dynamic mean field theory} The order parameters of the dynamic mean field theory are the correlation function \begin{equation} q(t-t')=\frac 1 N\sum_{\alpha}\overline{\,\av{\big[\varrho_{\alpha}(t) -\varrho_{\alpha}(t')\big]^2\,}\,} \label{q-def}\end{equation} and the response function \begin{equation} r(t-t')=\frac T N \sum_{\alpha}\overline{\,\delta\av{\varrho_{\alpha}(t)} /\delta F_{\alpha}(t')\Big.\,}\,. \label{r-def} \end{equation} They obey the mean field equations \cite{KibachHo93I,KibachHo93II} \begin{equation} \partial_t r(t)=-\mu\,r(t)+\int_0^t \d{s}w(s)r(s)r(t-s) \label{r-}\end{equation} and \begin{eqnarray} \partial_t q(t)&=&2-\mu\,q(t)+\int_0^t \d{s}w(s)r(s)q(t-s)\nonumber\\ &-&\int_0^{\infty}\d{s}\Big\{2\big[W(t+s)-W(s)\big]r(s)\nonumber\\ &-&\big[w(t+s)r(t+s)-w(s)r(s)\big]q(s)\Big\}\hspace{1.4cm} \label{q-}\end{eqnarray} with \begin{equation} w(t)=-4\beta^2f''\big(\,q(t)+v^2t^2\,\big)\,, \label{w-def}\end{equation} \begin{equation} W(t)=2\beta^2f'\big(\,q(t)+v^2t^2\,\big) \label{W-def}\end{equation} and \begin{equation} \mu=\mu_0+\int_0^{\infty}\d{s}w(s)r(s)\,. \label{mu-def}\end{equation} A brief derivation is given in the appendix. In the following $\mu_0=0$ is assumed. In thermal equilibrium correlation and response functions are related by a fluctuation-dissipation theorem (FDT) which reads $\partial_t q(t)=2r(t)$. This is violated in a non-equilibrium situation and \begin{equation} n(t)=\frac{1}{2r(t)}\,\partial_tq(t)-1 \label{n-def}\end{equation} is introduced as measure of this FDT-violation. Using this in \citeq{w-def} and \citeq{W-def} \begin{equation} \partial_t W(t)=-\Big\{\big[1+n(t)\big]r(t)+v^2t\Big\}w(t) \label{dW-}\end{equation} is found. Combining \citeq{r-}, \citeq{q-} and \citeq{n-def} one obtains \begin{eqnarray} r(t)\partial_tn(t)&=&-\int_0^t \d{s}w(s)r(s)r(t-s)\nonumber\\ &&\hskip20pt\times\big[n(t)-n(t-s)\big]\nonumber\\ &&+\int_0^{\infty}\d{s}\Big\{w(t+s)r(t+s)r(s)\nonumber\\ &&\hskip20pt\times\big[n(t+s)-n(s)\big]\nonumber\\ &&\hskip20pt+\,v^2\big(t+s\big)\,w(t+s)r(s)\Big\}\hspace{1.4cm} \label{n-}\end{eqnarray} and this equation can now be used instead of one of the original mean field equations \citeq{r-} or \citeq{q-}. Prescribing the drift velocity $v$ the average of the force necessary to drive the particle is \begin{equation} \beta \,F=v\,\Big\{1+\int_0^{\infty}\d{s}s\,w(s)r(s)\Big\} \label{F-}\end{equation} which is derived in the appendix. \subsection{Effective exponents} It is convenient to define \begin{equation} g(t)=t\,r(t) \label{g-}\end{equation} and then \citeq{n-def} is written as \begin{equation} t\partial_t q(t)=2\big\{1+n(t)\big\}g(t)\,. \label{dq-}\end{equation} In order to present the results of the numerical integration of the mean field equations effective time dependent exponents are introduced. They are also used in the evaluation of the convolution type integrals in the self-energies of the mean field equations \citeq{r-} and \citeq{n-}. The first two exponents are \begin{equation} \nu(t)=t\partial_t\ln g(t) \label{nu-def}\end{equation} and \begin{equation} \alpha(t)=-t\partial_t \ln w(t)\,. \label{alpha-def}\end{equation} With \begin{equation} k(t)=t\partial_t n(t) \label{k-def}\end{equation} the third exponent \begin{equation} \kappa(t)=t\partial_t\ln k(t) \label{kappa-def}\end{equation} is defined. \section{Numerical results} \setcounter{equation}{0} \subsection{Numerical procedure} The dynamic mean field equations to be solved consist of two integro-differential equations \citeq{r-}and \citeq{n-}, a first order differential equation \citeq{n-def}, the definition of $w(t)$, \citeq{w-def} and \citeq{f-def}, and the integral \citeq{mu-def} determining $\mu$. The second integral in \citeq{n-} and the integral \citeq{mu-def} extend overall times and therefore, evaluating the solution at some $t$ requires the knowledge of the solution at all times, not only at $s<t$. This means that the equations have to be iterated. This is the price one has to pay for investigating a steady state situation. A study of the relaxation from a non-equilibrium initial state on the other hand is free of this problem, but it requires to deal with functions depending on two time arguments. The wide span of time arguments ranging from $10^{-4}$ to $10^{36}$ requires a nonuniform discretization. The present calculation uses a homogeneous discretization of $\ln t$. The convolution type integrals in \citeq{r-} and \citeq{n-} can be evaluated by numerical integration using the values stored at the points of the above grid for one of the functions of the integrand and an interpolation for the other function. Alternatively, an approximative evaluation of the integrals using effective exponents, described in Section 4.4 is possible. For the actual calculation both methods have been combined. \mfigure{q-A} {Correlation function $q(t)$ in phase $\cal A$}{q-sr} \mfigure{q-B} {Correlation function $q(t)$ in phase $\cal B$}{q-lr} The mean field equations in the form written in Section 4 consist of several coupled nonlinear differential equations of first order. They can be integrated forward or backward in time. In addition they have to be iterated. The situation resembles in some sense the problem of solving \begin{equation} \frac {\;\d x(t)}{\;\d t}=f\big(x(t)\big)\,. \end{equation} Integrating forward in $t$ the solution typically approaches a fixed point $f(\bar x)=0$ with $f'(\bar x)<0$, whereas a fixed point with $f'(\bar x)>0$ is reached integrating backward in $t$. Selecting an appropriate initial value $x(t_0)$ somewhere between a fixed point with $f'(\bar x)>0$ and an adjacent one with $f'(\bar x)<0$, the solution can easily be found by integration in both directions. For the complete set of equations such fixed point situations show up at the crossover between the various scaling regimes discussed below. \subsection{Results} \mfigure{n-A} {$n(t)$ in phase $\cal A$}{n-sr} \mfigure{n-B} {$n(t)$ in phase $\cal B$}{n-lr} \mfigure{nu-A} {Effective exponent $\nu(t)$ in phase $\cal A$}{nu-sr} \mfigure{nu-B} {Effective exponent $\nu(t)$ in phase $\cal B$}{nu-lr} \mfigure{ka-A} {Effective exponent $\kappa(t)$ in phase $\cal A$}{ka-sr} \mfigure{ka-B} {Effective exponent $\kappa(t)$ in phase $\cal B$}{ka-lr} \mfigure{al-A} {Effective exponent $\alpha(t)$ in phase $\cal A$}{al-sr} \mfigure{al-B} {Effective exponent $\alpha(t)$ in phase $\cal B$}{al-lr} \mfigure{tv-A} {Characteristic time scales in phase $\cal A$}{tv-sr} \mfigure{tv-B} {Characteristic time scales in phase $\cal B$}{tv-lr} The following figures show selected results obtained for $\{\gamma=1.5;\; T=0.4\}$ and $\{\gamma=0.75;\; T=0.7\}$, respectively. The first set of variables corresponds to a situation where the replica calculation \cite{Mez:manifold90,Mez:manifold91} requires single step replica symmetry breaking. This corresponds to a point in phase $\cal A$ of the phase diagram, Fig.\ref{phase-fig}. The second set corresponds to a point in phase $\cal B$ and the corresponding replica treatment requires continuous replica symmetry breaking. The following figures show results obtained for various values of the drift velocity ranging from $v=10^{-6}$ to $v=10^{-30}$. In the following various velocity dependent characteristic time scales are introduced. Their values for $v=10^{-30}$ are marked in the figures. Figs.\ref{q-sr} and \ref{q-lr} show the correlation function $q(t)$. The plateau value $q_c$ is given by \citeq{T-FDT}. In view of the argument $q(t)+v^2t^2$ in \citeq{w-def} the value of $v^2t^2$ for $v=10^{-30}$ is also shown. The function $n(t)$ defined in \citeq{n-def} is shown in Figs.\ref{n-sr} and \ref{n-lr}. It indicates the violation of the FDT, which holds for $t<t_p$. It develops a first plateau value for $t_p<t<t_a$ and a second one for $t>t'_{a}$. In phase $\cal A$ this second value obviously depends on $v$. In phase $\cal B$ the first plateau is not yet fully developed, even for $v=10^{-30}$. The time dependent effective exponents $\nu(t)$, \citeq{nu-def}, $\kappa(t)$, \citeq{kappa-def}, and $\alpha(t)$, \citeq{alpha-def}, are shown in Figs.\ref{nu-sr} to \ref{al-lr}. Their properties are discussed below. From the data several characteristic time scales can be extracted. They are shown in Figs.\ref{tv-sr} and \ref{tv-lr} as functions of the velocity $v$. The time $t_p$ is defined by $q(t_p)=q_c$ and it characterizes the center of the plateau. The next time scale $t_x$ is relevant only for phase $\cal B$ and it is given by $\kappa(t_x)=0$. The time scale $t_a$ is defined by $q(t_a)=2\,q_c$ and $t_a'$ by $q(t_a')=2\,v^2{t_a'}2$. A power law dependence is found for $v\to0$. \subsection{Scaling regimes} The numerical results and analytic considerations described in Section 4 reveal the existence of various scaling regimes in the limit $v\to0$. \subsubsection{FDT-regime} For finite $t\sim t_0$ one finds $n(t)\ll1$ and fluctuation dissipation theorems hold. This can be understood as a situation where the particle stays within one valley of the energy landscape. The correlation function and $n(t)$ can be written as \begin{eqnarray} q(t)&=&\mathaccent"7617 q(t/t_0)\nonumber\\ n(t)&=&\mathaccent"7617 b(v)\,\mathaccent"7617 n(t/t_0) \label{scf-FDT}\end{eqnarray} with $\mathaccent"7617 b(v)\to0$ for $v\to0$. The time scale $t_0$ and the functions $\mathaccent"7617 q(x)$ and $\mathaccent"7617 n(x)$ do not depend on $v$. This regime is referred to as the FDT-regime. \subsubsection{Plateau-regime} With increasing $t$ a plateau regime is found where $q(t)\approx q_c$. The corresponding time scale is $t_p=t_p(v)$ defined by $q(t_p)=q_c$ and the above functions obey the scaling form \begin{eqnarray} q(t)&=&q_c+\hat a(v)\,\hat q(t/t_p)\nonumber\\ n(t)&=&n_c+\hat n(t/t_p) \label{scf-pl}\end{eqnarray} which is consistent with $\nu(t)=\hat\nu(t/t_p)$ and $\kappa(t)=\hat\kappa(t/t_p)$. The function $n(t)$, which is a measure of FDT-viola\-tion, reaches a plateau value $n_c$ for $t\sim t_p$. In phase $\cal A$ this plateau extends beyond the limits of the $q$-plateau. In phase $\cal B$ the plateau of $n(t)$ is not yet very pronounced, even at $v=10^{-30}$, but it can be seen that the upper limits of the $n$-plateau and the $q$-plateau essentially coincide. The center of the $n$-plateau defines an additional time scale $t_x=t_x(v)$ where $n(t_x)=n_c$ or $\kappa(t_x)=0$. For $t\sim t_x$ a scaling form \begin{equation} n(t)=n_c+c(v)\hat n_x\big(c(v)\ln t/t_x)\big) \label{n-x}\end{equation} will be derived. In phase $\cal C$ $\;n(t)\ll1$ holds up to the upper end of the $q$-plateau and consequently the FDT-solution holds in the whole plateau regime. \subsubsection{Asymptotic regime} The lower end of the asymptotic regime is marked by a time scale $t_a=t_a(v)$ where $q(t_a)-q_c\sim q_c$. An appropriate choice is $t_a$ such that $q(t_a)=2q_c$. For $t\sim t_a$ a scaling form \begin{eqnarray} q(t)&=&\bar q(t/t_a)\nonumber\\ n(t)&=&n_c+\bar b(v)\bar n(t/t_a) \label{scf-as}\end{eqnarray} with $\bar b(v)=1$ in phase $\cal B$ and $\cal C$ is found. Another time scale $t'_a= t'_a(v)$ can be defined such that $q(t'_a)=\big(v\,t'_a\big)^2$. In phase $\cal B$ and $\cal C$ both are proportional to $v^{-1}$, whereas $t_a(v)\sim v^{-1+\eta}$ and $t'_a(v)\sim v^{-1-\eta}$ with $\eta>0$ is found in phase $\cal A$. This requires for $t\sim t'_a$ the scaling form \begin{equation} n(t)=n_c+\bar b'(v)\,\bar n'(t/t_a ')\,. \label{scf-as'}\end{equation} The existence of the exponent $\eta>0$ is connected to the creep behavior observed in phase $\cal A$. The above results now have to be verified by analytic investigations. This is done in the following sections. \section{Analytic results} \setcounter{equation}{0} \subsection{FDT-solution} The shortest $v$-dependent time scale for $v\to0$ is $t_p$ with ${1\ll t_p\ll v^{-1}}$. For $t\ll t_p$ one expects a solution which obeys the FDT and therefore $n(t)=0$. This solution holds in phase $\cal O$ for $v=0$ and describes equilibrium within a single valley of the energy landscape in the other phases. The discussion of this solution follows standard arguments \cite{KibachHo93I}. With \citeq{q-}, \citeq{dW-} and \citeq{dq-} one obtains \begin{eqnarray} \partial_tq(t)&=&2-\mu q(t)\nonumber\\&&+\int_0^t\d{s}w(s)r(s)q(t-s) \label{dq-FDT}\,,\end{eqnarray} because the second integral in \citeq{q-} vanishes. In order to demonstrate this, one realizes that its main contributions come from $s\sim t$ where $n(t)\approx 0$. With \citeq{g-} and \citeq{dq-} $\partial_t q(t)\approx 2r(t)$ and with \citeq{dW-} $\partial_t W(t)\approx-w(t)r(t)$. This allows to write the integrand as a complete derivative with respect to $s$ and to evaluate the integral resulting in $0$, since $q(0)=0$ and $W(t)\to 0$ for $t\to\infty$. Eqs.\citeq{r-} and \citeq{dq-FDT} cannot be solved in closed form. It is, however, sufficient to investigate the solution for $t\to t_p$, where $q(t)\to q_c$, $\partial_t q(t)\to0$ and $\partial_t r(t)\to0$. In leading order the integrals in both equations can be evaluated by taking into account only the contributions near the upper and lower bound, respectively. This yields from \citeq{r-} \begin{equation} \mu=2\beta^2\big\{f'(0)-f'(q_c)-q_cf''(q_c)\big\} \label{mu-FDT}\end{equation} and from \citeq{dq-FDT} \begin{equation} f''(q_c)q_c^2=-T^2\,. \label{T-FDT}\end{equation} The second equation is used to determine $q_c$. Elimination of $\mu$ in \citeq{mu-FDT} using \citeq{mu-def} yields with \citeq{w-def},\citeq{W-def} and \citeq{dq-} \begin{eqnarray} q_cf''(q_c)&=&-\frac{1}{2\beta^2}\int_{t_p}^{\infty}\d{s}w(s)\,r(s)\nonumber\\ &=&\int_{t_p}^{\infty}\d{s}\dot q(s)\frac{f''\big(q(s)+v^2s^2\big)}{1+n(s)}\,. \label{condition}\end{eqnarray} This condition involving the solutions at long time scales will be used later. In order to give estimates of the corrections to the leading order one assumes \begin{equation} t\,r(t)=g(t)\mathop{\longrightarrow}_{t\to t_p} -\sfrac12\nu_0q_c\,(t/t_0)^{\nu_0} \label{g-FDT}\end{equation} with $\nu_0<0$. The time scale $t_0$ has yet to be determined. This leads to \begin{eqnarray} q(t)&&=q_c-2\int_t^{t_p}\d{s}r(s)\nonumber\\ &&\mathop{\longrightarrow}_{t\to t_p}q_c\Big\{1-(t/t_0)^{\nu_0}\Big\}\,. \label{q-FDT}\end{eqnarray} This is now used in an analysis of the corrections in \citeq{r-} where the leading order contains terms $\sim t^{\nu_0-1}$. The next to leading order $\sim t^{2\nu_0-1}$ results in \begin{eqnarray} &&f''(q_c)\int_0^t\d{s}\big\{r(s)r(t-s)-2r(t)r(s)\big\}\nonumber\\ &&\qquad\quad =\big\{f''(q_c)+\sfrac12 f'''(q_c)q_cr(t)\big\} \big\{q_c-q(t)\big\}\,.\qquad \end{eqnarray} With \citeq{g-FDT} the integral on the left hand side can be evaluated. Introducing \begin{equation} R_m=-\frac{q_c\,f'''(q_c)}{2f''(q_c)} \label{R-def}\end{equation} one obtains \begin{equation} \frac{\Gamma^2(1+\nu_0)}{\Gamma(1+2\nu_0)}=R_m \label{nuo-}\end{equation} which allows to determine the exponent $\nu_0$. In order to get an estimate of the time scale $t_0$ one can use an interpolation formula \begin{equation} q(t)\approx q_c\Big\{1-\big(1+t/t_0\big)^{\nu_0}\Big\} \end{equation} which gives the correct asymptotic behavior for $t\to t_p$. The requirement $q(t)\to 2t$ for $t\to0$ yields \begin{equation} t_0\approx-\sfrac12\nu_0q_c\,. \label{to-}\end{equation} In the ergodic high temperature phase $\cal O$ for $v=0$ the FDT holds for all times and therefore with \citeq{mu-def} \begin{equation} \mu=\frac{2f'(0)}{T^2}\,. \end{equation} At the transition to one of the nonergodic phases the plateau develops and one finds with \citeq{mu-FDT} and \citeq{T-FDT} \begin{eqnarray} T_c^2&=& q_cf'(q_c)\,,\nonumber\\ f'(q_c)&=&-q_cf''(q_c)\,. \label{Tc-}\end{eqnarray} A solution of these equations with finite $T_c$ exists only for $\gamma>1$. It is \begin{eqnarray} q_c&=&\frac1{\gamma-1}\nonumber\\ T_c&=&\sqrt{\sfrac12}\,\gamma^{-\gamma/2}\,(\gamma-1)^{(\gamma-1)/2}\,. \label{Tc--}\end{eqnarray} This is shown in Fig.\ref{phase-fig}. The corresponding value obtained from replica theory \cite{Mez:manifold90,Mez:manifold91} \begin{equation} T_{c,\rm1RSB}=\frac{1}{\sqrt{6\gamma}} \label{Tc-1RSB}\end{equation} is always below the above value. For $T>T_c$ in phase $\cal O$ the FDT holds for all times and the force \citeq{F-} is \begin{equation} \beta\,F=v\,\Big\{1+\int_0^{\infty}\d{t}W(t)\Big\}\,. \label{Foo}\end{equation} For $t\to \infty$ \citeq{r-} yields $r(t)\to r_{\infty}$ and $q(t)\sim t$. Therefore the integral in \citeq{Foo} converges for $\gamma>1$ and there exists a finite friction constant for $v\to 0$ \subsection{Integrals and counterterms} For $t\gg1$ the leading contributions to the integrals in \citeq{r-} and \citeq{n-} come from regions near the boundaries of integration. Subtracting those leads to residual integrals and a reformulation of the dynamic mean field equations. The following integrals are introduced: \begin{eqnarray} J(t)&=&\int_0^t\d{s}\Big\{w(s)r(s)r(t-s)-w(t)r(t)r(t-s)\nonumber\\ &&-w(s)r(s)r(t)+s\,w(s)r(s)\partial_t r(t)\Big\}\,,\quad\quad \label{J-def}\end{eqnarray} \begin{eqnarray} K(t)&=&\int_0^{\infty}\d{s}w(t+s)r(t+s)r(s)\nonumber\\ &&\quad\times\Big\{n(s)-n(s+t)\Big\}\nonumber\\ &&-\int_0^t\d{s}w(t-s)r(t-s)\Big\{r(s)n(s)\nonumber\\ &&\qquad-r(s)n(t)+(t-s)r(t)\partial_tn(t)\Big\}\qquad\quad \label{K-def}\end{eqnarray} and \begin{equation} U(t)=v^2\int_0^{\infty}\d{s}\big(t+s\big)w(t+s)r(s)\,. \label{U-def}\end{equation} Furthermore it is appropriate to define \begin{equation} D(t)=\int_t^{\infty}\d{s}w(s)r(s)-w(t)\int_0^t\d{s}r(s) \label{D-def}\end{equation} and \begin{equation} Z(t)=1+\int_0^t\d{s}\,s\,w(s)r(s)\,. \label{Z-def}\end{equation} This allows to rewrite the mean field equation \citeq{r-} in the form \begin{equation} Z(t)\,\partial_tr(t)=J(t)-D(t)r(t) \label{r-2}\end{equation} and \citeq{n-} as \begin{equation} Z(t)\,r(t)\partial_t n(t)=U(t)-K(t)\,. \label{n-2}\end{equation} \subsection{QFDT- and hierarchical solution} With \citeq{D-def} and $n(t)\approx0$ for $t<t_p$ the condition \citeq{condition} means $D(t_p)=0$. Depending on $\gamma$ for $v=0$ a QFDT-solution and a hierarchical solution, respectively, have been proposed \cite{KibachHo93I,KibachHo93II}. These solutions can easily be obtained from the above equations. Especially the second scheme requires, however, assumptions which are not fulfilled. Nevertheless they are presented here for further reference. \subsubsection{QFDT-solution} In the QFDT-solution $n(t)=n_Q$ for $t>t_p$ is assumed. From \citeq{condition} one finds \begin{equation} n_Q+1=-\frac{f'(q_c)}{q_c f''(q_c)}\,. \label{nQ-def}\end{equation} Without further analysis of the regime $t>t_p$ nothing can be said about the range of validity of this solution or about the force $F$. This solution was proposed to be valid for $\gamma>1$ and it shows some similarity to the 1RSB-calculation \cite{Mez:manifold90,Mez:manifold91}. \subsubsection{Hierarchical solution} Assume that the integral in \citeq{r-} for $t>t_p$ is completely determined by its contributions from the upper and lower bound, respectively. This means that $J(t)\approx 0$. As it turns out this is the essence of the proposal of the existence of an ultrametric organization of long time scales. Neglecting the derivative with respect to $t$ in \citeq{r-2} results in $D(t)\approx 0$ for $t>t_p$. Defining \begin{equation} m\big(q(t)\big)=\frac{1}{1+n(t)} \end{equation} Eq.\citeq{D-def} yields \begin{equation} \int_q^{\infty}\d{q'}f''(q')m(q')=f''(q)\int_0^q\d{q'}m(q')\,. \label{m-hierarch}\end{equation} This leads to the differential equation \begin{equation} \partial_q \ln m(q)=\frac{f''''(q)}{f'''(q)}-\frac{3f'''(q)}{2f''(q)} \end{equation} which is solved for $q\ge q_c$ by \begin{equation} m(q)=m(q_c)\left\{\frac{1+q}{1+q_c}\right\}^{-\sfrac12(1-\gamma)} \end{equation} using \citeq{f-def}. The integration constant $m(q_c)$ is obtained from \linebreak \citeq{m-hierarch} with $q=q_c$ \begin{equation} m(q_c)=-\frac{f'''(q_c)q_c}{2\,f''(q_c)}=R_m \end{equation} with $R_m$ given in \citeq{R-def}. This is the solution obtained previously, assuming a hierarchical structure of long time scales \cite{KibachHo93II}. \subsection{Evaluation of integrals via effective exponents} The main problem in a discussion of the solutions of the mean field equations and also in their numerical integration is in the evaluation of the integrals listed in the previous section. In the following an approximative scheme is proposed, which turns out to be very accurate over the whole range of $t$. This scheme is based on the effective time dependent exponents, which have been introduced in Section 2.3. It allows to write and evaluate the integrals in the form \begin{equation} \int_0^1\d{x}x^{a-1}(1-x)^{b-1}=\frac{\Gamma(a)\Gamma(b)}{\Gamma(a+b)}\,. \label{e-integral}\end{equation} As a first step in this program one introduces "dimen\-sion\-less" quantities instead of the original ones given in \citeq{J-def} to \citeq{Z-def}: \begin{equation} \tilde J(t)=\frac{t\,J(t)}{w(t)\,g^2(t)}\,, \label{J-.}\end{equation} \begin{equation} \tilde K(t)=\frac{t\,K(t)}{k(t)\,w(t)\,g^2(t)}\,, \label{K-.}\end{equation} \begin{equation} \tilde U(t)=\frac{t\,U(t)}{k(t)\,w(t)\,g^2(t)}\,, \label{U-.}\end{equation} \begin{equation} \tilde D(t)=\frac{D(t)}{w(t)\,g(t)}\,, \label{D-.}\end{equation} and \begin{equation} \tilde Z(t)=\frac{Z(t)}{t\,w(t)\,g(t)}\,. \label{Z-.}\end{equation} The functions in the integrands are approximated by \begin{equation} s\,r(s)=g(s)\approx g(t)\left(\frac s t\right)^{\nu(t)}\,, \label{r(s)-}\end{equation} \begin{equation} w(s)\approx w(t)\left(\frac s t\right)^{-\alpha(t)} \label{w(s)-}\end{equation} and \begin{equation} n(s)\approx n(t)+\frac{k(t)}{\kappa(t)} \left\{\left(\frac s t\right)^{\kappa(t)}-1\right\}\,. \end{equation} This ansatz fulfills \citeq{nu-def}, \citeq{alpha-def}, \citeq{k-def} and \citeq{kappa-def} for $s=t$. The integrals $\tilde J(t)$ and $\tilde K(t)$ are now of the form proposed above (for the first integral in \citeq{K-def} a trivial substitution of variables is necessary) and can be evaluated using \citeq{e-integral} provided $a>0$ and $b>0$. Otherwise counterterms are required cancelling the poles of the $\Gamma$-functions. The last three terms in \citeq{J-def} are of this kind. Depending on the values of the effective exponents further counterterms might be required. This yields (in the following the arguments $t$ are dropped) \begin{eqnarray} \tilde J&\approx &\frac{\Gamma(\nu-\alpha)\Gamma(\nu)}{\Gamma(2\nu-\alpha)}\nonumber\\ &&-\frac 1 {\nu}-\frac 1 {\nu-\alpha}+\frac{\nu-1}{\nu+1-\alpha}\qquad\quad \label{J-.-}\end{eqnarray} and \begin{eqnarray} \tilde K&\approx& \frac{1}{\kappa} \bigg\{\frac{\Gamma(\nu+\kappa)\Gamma(1+\alpha-2\nu-\kappa)} {\Gamma(1+\alpha-\nu)}\nonumber\\ &&\qquad-\frac{\Gamma(\nu)\Gamma(1+\alpha-2\nu-\kappa)} {\Gamma(1+\alpha-\nu-\kappa)}\nonumber\\ &&\qquad+\frac{\Gamma(\nu-\alpha)\Gamma(\nu)} {\Gamma(2\nu-\alpha)}-\frac{\Gamma(\nu-\alpha)\Gamma(\nu+\kappa)} {\Gamma(2\nu+\kappa-\alpha)}\nonumber\\ &&\qquad-\frac{\kappa}{\nu-\alpha+1}\bigg\}\,. \label{K-.-}\end{eqnarray} The integral \citeq{U-.} is given for $\nu>0$ and $\alpha-\nu>1$ by \begin{equation} \tilde U\approx \frac{v^2\,t^2}{k\,g}\,\frac{\Gamma(\nu)\Gamma(\alpha-\nu-1)} {\Gamma(\alpha-1)} \label{U-.-1}\end{equation} otherwise it is approximated by \begin{equation} \tilde U\approx \frac{v^2\,t\,Z(\infty)}{k\,w\,g^2}\,. \label{U-.-2}\end{equation} For $\alpha-\nu>1$ \citeq{Z-.} becomes \begin{equation} \tilde Z\approx \frac{Z(\infty)}{t\,w\,g} \label{Z-.-2}\end{equation} otherwise \begin{equation} \tilde Z\approx \frac 1 {1+\nu-\alpha}\,. \label{Z-.-1}\end{equation} The mean field equations \citeq{r-2} and \citeq{n-2} are rewritten as \begin{equation} \big\{\nu-1\big\}\,\tilde Z=\tilde J-\tilde D \label{r-.-}\end{equation} and \begin{equation} \tilde Z=\tilde U-\tilde K\,. \label{n-.-}\end{equation} The exponent $\alpha$, \citeq{alpha-def}, is \begin{equation} \alpha=-2\,g\,\big\{1+n+u\big\}\frac{f'''(q+v^2t^2)}{f''(q+v^2t^2)} \label{alpha-.-}\end{equation} with \begin{equation} u=\frac{v^2\,t^2}{g}\,. \label{u-def}\end{equation} Differentiation of \citeq{D-def} yields with \citeq{D-.} \begin{equation} t\partial_t \tilde D=\big\{\alpha-\nu\big\}\tilde D-2\big\{1-[1+n+u]\,R \big\} \label{dD-.-}\end{equation} with \begin{equation} R=-\frac{f'''(q+v^2t^2)}{f''(q+v^2t^2)}\int_0^t\d{s}r(s)\,. \label{R-.-}\end{equation} The complete set of mean field equations now involves 15 functions of time: $q$, $r$, $g$, $n$, $k$, $\nu$, $\kappa$, $\alpha$, $u$, $R$, $\tilde D$, $\tilde J$, $\tilde K$, $\tilde U$ and $\tilde Z$. The corresponding 15 equations are: \citeq{g-}, \citeq{dq-}, \citeq{nu-def}, \citeq{k-def}, \citeq{kappa-def}, \citeq{J-.-}, \citeq{K-.-}, (\ref{U-.-1}/\ref{U-.-2}), (\ref{Z-.-2}/\ref{Z-.-1}), \citeq{r-.-}, \citeq{n-.-}, \citeq{u-def}, \citeq{alpha-.-}, \citeq{dD-.-} and \citeq{R-.-}. Some of the variables can easily be eliminated, but even then one is left with 5 differential equations of first order and 2 implicit ordinary equations for 7 of the above functions of time. This certainly looks complicated. On the other hand, in each of the scaling regimes to be discussed below only a subset of variables and equations has to be looked at. \subsection{The plateau regime} Having discussed the FDT-solution in Section 4.1 already, the investigation of the plateau regime follows next. This regime is characterized by $\big|q(t)-q_c\big|\ll q_c$. Further simplifications are due to $t^2 v^2\ll q_c$. Consequently $R$ defined in \citeq{R-.-} can be replaced by the constant $R_m$ given in \citeq{R-def}. Furthermore $\tilde U\approx0$ and $u\approx0$ can be used. From \citeq{alpha-def} with \citeq{w-def} one finds $\alpha\approx0$ and then \citeq{Z-.-1} yields $\tilde Z\approx\frac 1 {1+\nu}$. Eqs.\citeq{J-.-}, \citeq{r-.-} and \citeq{Z-.-1} determine $\nu=\tilde\nu(\tilde D)$ as a function of $\tilde D$, shown in Fig.\ref{nu(D)}. \mfigure{nu-D} {$\nu(\tilde D)$ and $-\tilde D\nu(\tilde D)$ as functions of $\tilde D$, see text.}{nu(D)} Eqs.\citeq{K-.-} and \citeq{n-.-} are solved by \begin{equation} \kappa(t)=1-3\nu(t)\,. \label{ka3nu}\end{equation} Eq. \citeq{dD-.-} now reads \begin{eqnarray} t\partial_t \tilde D(t)&=&-\tilde\nu\big(\tilde D(t)\big)\tilde D(t)\nonumber\\ &&-2\big\{1-\big[1+n(t)\big]R_m\Big\}\,. \label{dD-lin}\end{eqnarray} The remaining set of equations to be solved consists of \citeq{k-def}, \citeq{kappa-def}, \citeq{J-.-}, \citeq{Z-.-1}, \citeq{r-.-}, \citeq{ka3nu} and \citeq{dD-lin}. Time enters only via derivatives of the form $t\partial_t$ and therefore a family of scale invariant solutions exists. This means that the whole family can be written as a set of functions depending only on $t/t_p$ with arbitrary $t_p$. This is in accordance with the scaling form proposed in \citeq{scf-pl}. The time scale $t_p\gg t_0$ can now be fixed such that $\tilde D(t_p)=0$ and as a consequence $\nu(t_p)=0$ and $\kappa(t_p)=1$. Integrating \citeq{dD-lin} down to $t\ll t_p$ a stable fixpoint $\tilde D(t)\to \tilde D_0>0$ is reached with \begin{equation} -\tilde\nu(\tilde D_0)\tilde D_0=2\big\{1-R_m\big\} \end{equation} and $n(t)\to 0$. Eq.\citeq{nuo-} yields $\nu(\tilde D_0)=\nu_0$. For $t_0\ll t \ll t_p$ the function $q(t)$ can be matched to the asymptotic FDT-solution \citeq{q-FDT} and \begin{equation} \hat a(v)=\big(t_p/t_0\big)^{\nu_0} \label{ahat}\end{equation} results for the scale factor in \citeq{scf-pl} where $t_p=t_p(v)$ has to be determined later by matching to the asymptotic regime. It can be shown that the above fixing of $t_p$ is equivalent to the original definition $q(t_p)=q_c$ within a factor close to $1$. The discussion of the asymptotes for $t\gg t_p$, still within $\big|q(t)-q_c\big|\ll q_c$, is more involved. Neglecting the derivative with respect to time in \citeq{dD-lin} one gets \begin{equation} -\tilde D\,\tilde\nu(\tilde D)=2\big\{1-(1+n)R_m\big\} \end{equation} which now determines $\tilde D=\tilde D_1(n)$, $\nu_1=\tilde\nu\big(\tilde D_1(n)\big)$ and $\kappa_1=\tilde\kappa\big(\tilde D_1(n)\big)=1-3\tilde\nu\big(\tilde D_1(n)\big)$. There is a special value $n_x$ such that $\tilde\kappa\big(\tilde D_1(n_x)\big)=0$ with $\tilde D_1(n_x)=\tilde D_x\approx-0.700084$. This value is found from \citeq{dD-lin} as \begin{equation} n_x=\frac{1+\frac 1 6 \tilde D_x}{R_m}-1\,. \label{nx-def}\end{equation} Expanding \begin{equation} \tilde\kappa\big(\tilde D_1(n)\big)\approx \big(n-n_x)\,\kappa' \end{equation} with $\kappa'>0$, Eqs.\citeq{k-def} and \citeq{kappa-def} can be integrated in closed form. The real solutions are \begin{equation} n(t)=n_x-\frac{2\,c}{\kappa'}\,\coth\big(c\,\ln(t/t_x)\big) \label{n1-sr}\end{equation} and \begin{equation} n(t)=n_x+\frac{2\,c}{\kappa'}\,\tan\big(c\,\ln(t/t_x)\big) \label{n1-lr}\end{equation} with constants of integration $c$ and $t_x$. There are now two characteristic values for $n(t)$, the value $n_Q$ defined in \citeq{nQ-def} and $n_x$ defined above in \citeq{nx-def}. Both are functions of $T$ and $\gamma$. The relative magnitude of $n_Q$ and $n_x$ determines which of the above solutions has to be selected, as will be discussed in the following. This, on the other hand, determines ultimately whether creep or pinning is observed and therefore the different phases are characterized by their values of $n_Q$ and $n_x$. Phase $\cal A$ is defined as the region where $n_x>n_Q>0$, phase $\cal B$ has $n_Q>n_x>0$ and $n_Q>0>n_x$ holds in phase $\cal C$. The ergodic phase $\cal O$ with finite friction has already been discussed in Section 4.1 \subsubsection{Phase $\cal A$} The numerical results shown in Fig.\ref{n-sr} indicate a plateau of $n(t)$ extending to times beyond the upper boundary of the $q$-plateau. This suggests that the solution \citeq{n1-sr} has to be used for $t>t_p$. The choice $t_x\sim t_p$ yields the scaling form \citeq{scf-pl}. For $t\gg t_p$ the requirement $n(t)\to n_Q<n_x$ determines \begin{equation} c=\frac{\kappa'}{2}\,\big(n_x-n_Q\big)\approx - \frac{\kappa_1}{2} \end{equation} with $\kappa_1=\tilde\kappa\big(\tilde D_1(n_Q)\big)$. The parameter $n_Q$ is later shown to agree with the value obtained for the QFDT-solution \citeq{nQ-def}. For $t\gg t_p$ the above equations yield with $\nu_1=\tilde\nu\big(\tilde D_1(n_Q)\big)>0$ and $\kappa_1=1-3\nu_1<0$ \begin{eqnarray} q(t)&\to&q_c+\big(t_p/t_0\big)^{\nu_0}\,\big(t/t_p\big)^{\nu_1}\,\hat q_1\nonumber\\ n(t)&\to&n_Q-\big(t/t_p\big)^{\kappa_1}\,\hat n_1\,. \label{qn-A}\end{eqnarray} For $v\to0$ the constants $\hat q_1$ and $\hat n_1$ depend only on $\gamma$ and $T$. For $t\ll t_p$ \begin{equation} n(t)\to \hat n_0\big(t/t_p\big)^{\kappa_0} \label{n0-A}\end{equation} with constant $\hat n_0$ and $\kappa_0>0$. Matching at $t_0\ll t\ll t_p$ yields the scale factor \begin{equation} \mathaccent"7617 b(v)=\big(t_0/t_p\big)^{\kappa_0} \label{b0-match}\end{equation} in \citeq{scf-FDT}. The time scale $t_p=t_p(v)$ is still open and has to be determined later by matching to the asymptotic regime. The same holds for $n_Q$. \subsubsection{Phase $\cal B$} The plateau of $n(t)$ in this phase extends only to the upper boundary of the $q$-plateau, as shown in Fig.\ref{n-lr}. This indicates that solution \citeq{n1-lr} now has to be used for $t>t_p$. Its range of validity is restricted to $-\sfrac12\pi<c\ln(t/t_x)<\sfrac12\pi$ or $\mbox{\rm e}^{-\pi/2c}t_x<t<\mbox{\rm e}^{\pi/2c}t_x$. Later it is shown that this range increases for $v\to0$ and therefore $c=c(v)\to0$ for $v\to0$. For $t\to t_p$ the solution should not depend on $c$, which is the case for the choice \begin{equation} t_x=\mbox{\rm e}^{\pi/2c}\,t_p\,. \label{tx-B}\end{equation} For $t_p\ll t\ll \mbox{\rm e}^{\pi/c}\,t_p$ the second term in \citeq{n1-lr} is a small correction and the proposed scaling form \citeq{scf-pl} and \citeq{n-x} holds over the whole range $t_0\ll t\ll \mbox{\rm e}^{\pi/c}\,t_p$. With $\kappa_1=\tilde\kappa\big(\tilde D_1(n_x)\big)=0$ and $\nu_1=\tilde\nu\big(\tilde D_1(n_x)\big)=\frac13$, \ $q(t)$ has again the asymptotic form \citeq{qn-A} for $t\gg t_p$ and the scale factor $\mathaccent"7617 b(v)$ is given by \citeq{b0-match}. In this phase the quantities to be determined later by matching to the asymptotic regime are $t_p=t_p(v)$ and $t_x=t_x(v)$. \subsubsection{Phase $\cal C$} In this phase with $n_x<0$ again solution \citeq{n1-sr} has to be used. Since now $\kappa_1=\tilde\kappa\big(\tilde D_1(n_x)\big)>0$ and $n(t_p)\ll1$, the appropriate choice of the parameters is \begin{equation} \frac{2c}{\kappa'}= - n_x\,;\qquad c=\sfrac12\kappa_1\,. \end{equation} This yields for $t_x\gg t_p$ and $t_p\ll t\ll t_x$ \begin{equation} n(t)\approx -2 \, n_x\,\big(t/t_x\big)^{\kappa_1}\ll1 \label{n1-C}\end{equation} which means that \citeq{scf-pl} is fulfilled with $n(t)\approx0$. The asymptotic form of $q(t)$ is again \citeq{qn-A} with $\nu_1=\tilde\nu\big(\tilde D_1(0)\big)$. Again the quantities to be determined later by matching to the asymptotic regime are $t_p=t_p(v)$ and $t_x=t_x(v)$. \subsection{The asymptotic regime} The discussion so far did not depend on the actual choice of the drift velocity $v$. On the other hand, there are parameters not determined yet. These are the time scale $t_p(v)$ for all phases and the second time scale $t_x(v)$ for phase $\cal B$ and $\cal C$. For phase $\cal A$ it has to be shown that $n_Q$ actually agrees with the value obtained in the QFDT-solution, \citeq{nQ-def}. The velocity $v$ enters the full set of mean field equations in $\tilde U(t)$, (\ref{U-.-1}/\ref{U-.-2}), at various places in the argument of $f\big(q(t)+v^2t^2\big)$ and its derivatives, and in the definition of $u(t)$, \citeq{u-def}. In order to obtain the scaling form proposed in \citeq{scf-as} for $t\sim t_a$ it is necessary that $\tilde U(t)$ can be written as $\tilde U(t)=\bar U(t/t_a)$, which is the case for \begin{equation} \frac{v^2\,t_a^2(v)}{\bar b(v)}=1 \label{U-match}\end{equation} assuming the proposed scaling form in (\ref{U-.-1}/\ref{U-.-2}). For $t\ll t_a$ the solution in the asymptotic regime has to match the solution valid in the plateau regime obtained in the preceding section for $t\gg t_p$. This means \begin{equation} q(t)\mathop{\longrightarrow}_{t\to t_p} q_c+\big(t/t_a\big)^{\nu_1}\,\bar q_0 \end{equation} with constant $\bar q_0$. Comparison with \citeq{qn-A} yields \begin{equation} \big(t_p/t_0\big)^{\nu_0}\,t_p^{-\nu_1}\sim t_a^{-\nu_1}\,;\qquad t_p\sim t_0^{1-\zeta}\,t_a^\zeta \label{ta-tp}\end{equation} with \begin{equation} \zeta=\frac{\nu_1}{\nu_1-\nu_0}\,. \label{zeta}\end{equation} The discussion of $n(t)$ has to be done for the different phases separately. \subsubsection{Phase $\cal A$} Following \citeq{scf-as} one can write \begin{equation} n(t)\to n_Q-\big(t/t_a\big)^{\kappa_1}\,\bar b\,\bar n_0 \label{n-as}\end{equation} with constant $\bar n_0$ and $\bar b=\bar b(v)$. Comparison with \citeq{qn-A} results in \begin{equation} \bar b\sim \big(t_a/t_p)^{\kappa_1}\,. \label{bbar-A}\end{equation} Using this in \citeq{U-match} one finds with \citeq{ta-tp} \begin{eqnarray} t_a&\sim&v^{-1+\eta}\,t_0^{\eta}\nonumber\\ t_p&\sim&v^{-(1-\eta)\zeta}\,t_0^{1-(1-\eta)\zeta} \label{tv-A}\end{eqnarray} with \begin{equation} \eta=\frac{\nu_0\,\kappa_1}{2(\nu_1-\nu_0)+\nu_0\kappa_1}\,. \label{eta-def}\end{equation} Since $\nu_1>0$, $\nu_0<0$ and $\kappa_1<0$ one finds $\eta>0$. For $t\sim t_a$ the remaining $v$-dependence in the mean field equations is not of relevance, since \begin{equation} v^2t^2=(v\,t_0)^{2\eta}(t/t_a)^2\ll \bar q(t/t_a) \end{equation} and \begin{equation} u(t)=(v\,t_0)^{2\eta}(t/t_a)^2/\bar g(t/t_a)\ll 1 \end{equation} with $u(t)$ defined in \citeq{u-def} and $\bar g(t/t_a)=g(t)$. For $t\gg t_a$ one obtains $\nu(t)\to1$, $ r(t)\to \bar r_1/t_a$ and $q(t)\to \bar q_1\,t/t_a$ where $\bar r_1$ and $\bar q_1$ are constants. This follows immediately from \citeq{Z-.-2} and \citeq{r-.-} realizing that $\tilde Z(t)$ diverges $\sim t^{\alpha-2}$, whereas the right hand side of \citeq{r-.-} remains finite for $t\gg t_a$. This defines an even longer time scale \begin{equation} t_a'\sim v^{-1-\eta}\,t_0^{-\eta}\, \end{equation} where $q(t_a')=(vt_a')^2$. For $t_a\ll t\ll t_a'$ Eq. \citeq{alpha-def} yields $\alpha\to \gamma+1$. Investigating Eq.\citeq{n-.-} one observes that $\tilde Z(t)$, \citeq{Z-.-2}, diverges $\sim t^{\alpha-2}$ and $\tilde U(t)$, \citeq{U-.-1}, $\sim t^{1-\kappa}$, whereas $\tilde K(t)$, \citeq{K-.-} remains finite. This means $\kappa(t)\approx 3-\alpha(t)$ and $\kappa(t)\to \kappa_2=2-\gamma$. For $\gamma<2$ the exponent $\kappa(t)$ increases with increasing $t$ reaching a value $\kappa_2>0$ and therefore $n(t)$ also starts to increase again for $t_a\ll t\ll t_a'$ reaching a new constant value $n_2(v)$ for $t\gg t_a'$. Matching with \citeq{n-as} and \citeq{bbar-A} yields \begin{equation} n(t)=n_Q+\big(v\,t_0\big)^{2\eta(1-\kappa_2)}\,\bar n'(t/t_a') \end{equation} and therefore $\displaystyle n_2(v)-n_Q\sim \big(v\,t_0\big)^{2\eta(1-\kappa_2)}\mathop{\longrightarrow}_{v\to0}0$. For $\gamma>2$ this second plateau of $n(t)$ is missing, the conclusions are, however, unchanged. In order to give an estimate of the driving force one has to investigate the behavior of $\alpha(t)$, Eq.\citeq{alpha-def}, first. In the plateau region $\alpha(t_p)\approx 0$ was found. Around $t\sim t_a$ it starts to increase and reaches a value $\alpha(t)\approx\gamma+1$ for $t_a \ll t \ll t_a'$. Around $t\sim t_a'$ it starts to increase again reaching its asymptotic value $\alpha(t)=2(\gamma+1)$ for $t\gg t_a'$. The integral in the expression \citeq{F-} for the driving force gets its main contribution from $s$ such that $1+\nu(s)-\alpha(s)=0$. For $\gamma>1$ this is fulfilled for $s\sim t_a$ and $n(t)$ is well approximated by $n_Q$. This yields the following relation between mean velocity and driving force: \begin{equation} F=(v\,t_0)^{\eta}\,\bar F \end{equation} with \begin{equation} \bar F=-\frac{2\beta}{1+n_Q}\int_0^{\infty}\d x\,x\,f''\big(\bar q(x)\big)\,\partial_x \bar q(x) \end{equation} which does not depend on $v$. For given force the velocity increases slowly with increasing force according to \begin{equation} v\sim F^{1/\eta}\,. \label{v-creep}\end{equation} This behavior is a form of creep. In the one-dimensional case \cite{leDou95} creep has also been found for $\sfrac12<\gamma<1$ with \begin{equation} v\sim \mbox{\rm e}^{a\,F^{\mu}} \label{v(F)-1}\end{equation} with $\mu=2(\gamma-1)/(2\gamma-1)$. A power law dependence of the above form \citeq{v-creep} is obtained for $\gamma=1$. Note that a different definition of $\gamma$ is used in \cite{leDou95}. The results quoted here refer to the present definition \citeq{f-def}. The parameter $n_Q$ can be determined from the condition \citeq{condition} which is equivalent to the requirement $D(t_p)=0$. The main contribution to the integral comes from the region where $\nu(s)-\alpha(s)=0$ which is again the case for $s\sim t_a$. With $n(s)\approx n_Q$ the result \citeq{nQ-def} is recovered and $n_Q$ indeed agrees with the value derived within the QFDT-solution. It is remarkable that the longest time scale in phase $\cal A$ is not the external time scale $v^{-1}$ but rather $t'_a\sim v^{-1-\eta}$, which is longer. The ultimate reason for that is the behavior of $\kappa(t)=\kappa_1<0$ at the border between the plateau and the asymptotic regime. This value also enters the exponent $\eta$, \citeq{eta-def}. \subsubsection{Phase $\cal B$} This is no longer the case in phase $\cal B$. The choice \begin{equation} t_a\sim v^{-1} \label{ta-B}\end{equation} allows to rewrite the complete set of mean field equations for $t\sim t_a$ in terms of functions of $t/t_a$ only, especially \begin{equation} n(t)=\bar n(t/t_a)\,. \label{n-B}\end{equation} In the plateau region $n(t)$ is given by \citeq{n1-lr} and the dependence on $c$ drops out for $c\ln(t/t_x)\sim\pi/2$, or with \citeq{tx-B} for $t\sim \mbox{\rm e}^{\pi/c}t_p$. Matching with \citeq{n-B} yields \begin{equation} t_a\sim \mbox{\rm e}^{\pi/c}\,t_p\,. \label{ta-tp-c}\end{equation} On the other hand, \citeq{ta-tp} results from matching $q(t)$ and therefore \begin{equation} c=\frac{\pi}{(\zeta-1)\ln(t_0v)}\,. \end{equation} This means $c=c(v)\to0$ for $v\to0$ as proposed earlier. With \citeq{tx-B} the intermediate time scale $t_x$ is \begin{equation} t_x\sim t_0^{(1-\zeta)/2}\,v^{-(1+\zeta)/2}\,. \end{equation} The force necessary to sustain the velocity $v$ has to be calculated from \citeq{F-}. The main contribution comes as before from the region where $1+\nu(s)-\alpha(s)=0$, which is the case for $s\sim t_a$. Using the scaling form $q(t)=\bar q(t/t_a)$, \citeq{scf-as}, and \citeq{n-B} yields \begin{eqnarray} F&=&F_p\nonumber\\ &=&-2\beta\int_0^{\infty}\d x\,x\,f''(\bar q(x)+x^2)\frac{\partial_x\bar q(x)} {1+\bar n(x)}\qquad \label{Fp-BC}\end{eqnarray} This does not depend on $v$ and therefore a finite pinning force $F_p$ exists, which has to be overcome in order to set the particle in motion. \subsubsection{Phase $\cal C$} In phase $\cal C$ again \begin{equation} t_a=v^{-1} \end{equation} has to be chosen and matching to \citeq{n1-C} requires $t_x=t_a$. Otherwise the same arguments as above (phase $\cal B$) hold and there is again a finite pinning force given by \citeq{Fp-BC}. The main difference between phase $\cal B$ and $\cal C$ is the value of $n(t)$ for $t_p\ll t\ll t_a$, which is finite in phase $\cal B$ and essentially zero in phase $\cal C$. It is again of interest to compare this result with the one-dimensional case \cite{Sinai82,leDou95} where a finite pinning force is found for $\gamma=\sfrac12$, whereas the pinning force diverges for $\gamma<\sfrac12$. This divergence is due to the fact that no short distance cutoff in the correlation of the disorder is used in the one-dimensional calculation. This is also the reason why the phase boundaries in this case do not depend on temperature. \subsection{The phase diagram} Let me summarize the results regarding the phase diagram and the dependence of $v(F)$. Depending on temperature $T$ and exponent $\gamma$ several phases have been found. The above considerations yield the phase diagram shown in Fig.\ref{phase-fig} and discussed in more detail below. Different dependencies of the velocity on the driving force are observed in different phases. Examples are shown in Fig.\ref{v(F)}. \mfigure{v-F} {Drift velocity as a function of the applied force in phase ${\cal O}$, ${\cal A}$, ${\cal B}$ and at the boundary between phase ${\cal O}$ and ${\cal A}$.} {v(F)} The curve marked .75/.7 belongs to $\gamma=0.75$ and $T=0.7$, which is in phase $\cal B$. This point is also used for the examples presented in Section 3.2. Pinning is clearly observed. With increasing force $v(F)$ becomes linear in $F$ indicating a finite mobility $m=v/F$. Its value is determined by the first term in \citeq{F-}, which means that the influence of the random potential vanishes for high velocities. The curves marked 0.40 and 0.42 correspond to $\gamma=1.5$ and $T=0.4$ and $T=0.42$, respectively. Both points are in phase $A$. The temperature $T=0.4$ is the second example in Section 3.2. A power law dependence of $v(F)$ is found in accordance with the analytic investigations \citeq{v-creep}. For $\gamma=1.5$ and $T=0.4$ one finds $\eta=0.053$. With increasing force again the free mobility is found. The curve marked 0.44 is right at the transition line $\gamma=1.5$, $T=T_c(1.5)=0.4387$, and $\eta=0.19$ is found. The remaining curves marked 0.46 and 0.48 are in the high temperature phase $\cal O$ with $T=0.46$ and $T-0.48$, respectively. For $v\to0$ a finite mobility is found, which is much smaller than the free value. It vanishes at the critical temperature $T_c$. \subsubsection{Drift phase $\cal O$} For $T>T_c(\gamma)$, \citeq{Tc--}, and $\gamma>1$ a finite mobility $m=v/F$ is found for $v\to0$. It is given by \citeq{Foo} and it vanishes at the boundaries of this phase. There are no long time scales. \subsubsection{Creep phase $\cal A$} This phase is characterized by $n_Q<n_x$ given in \citeq{nQ-def} and \citeq{nx-def}. Its boundary with phase $\cal O$ is $T=T_c(\gamma)$ and the boundary to phase $\cal B$ is determined by $n_Q=n_x$. This yields with \citeq{R-def}, \citeq{nQ-def} and \citeq{nx-def} \begin{equation} \frac{f'(q_c)\,f'''(q_c)}{f''^2(q_c)}=2\,(1+\mbox{$\frac16$}\tilde D_x)\,. \end{equation} Inserting \citeq{f-def} one finds that this phase boundary is determined by $\gamma=\gamma_c$, with \begin{equation} \gamma_c=\frac{1}{1+\frac13\tilde D_x}\approx1.3044\,. \label{gammac-}\end{equation} For ${v\to 0}$ one observes creep in the form ${v\sim F^{1/\eta}}$, \citeq{v-creep}, with $\eta\to0$ for $\gamma\to\gamma_c$. \subsubsection{Pinning phase $\cal B$ and $\cal C$} This phase exists for $\gamma<1$ or $T<T_c$ and $\gamma<\gamma_c$. It has a finite pinning force which has to be overcome in order to set the particle in motion. Phase $\cal B$ and $\cal C$ are distinct only by their value of $n(t)$ for $t_p\ll t\ll t_a$. In phase $\cal C$ one finds $n(t)\approx0$ for $ t\ll t_a$ which means that the FDT holds up to this value, whereas it starts to be violated already for $t\sim t_a$ in phase $\cal A$ and $\cal B$. \section{Discussion} \setcounter{equation}{0} The first aspect of this investigation deals with the motion of a particle in a correlated random potential with power law decay of the correlations under the influence of an applied driving force. The phase diagram shows a phase with finite mobility, a creep phase and pinning phases. Similar behavior is found in a one-dimensional model \cite{Sinai82,leDou95} indicating that the transitions found are not an artifact of the mean field treatment, which becomes exact in the limit of infinite dimensionality. In the creep phase $v(F)$ obeys a power law. Such a behavior is also found in the one-dimensional case, but only at the boundary between the drift and the creep phase, which is otherwise ruled by a stretched exponential law. The pinning phase in the present case has a finite pinning force, which is also found in the one-dimensional case, but only at the boundary between creep and pinning phase. Otherwise the pinning force diverges in this calculation, which can be traced back to the absence of a short distance cutoff of the power law decay of the correlations. The numerical results indicate the existence of several scaling regimes which are verified by analytic investigations of the asymptotic properties in the limit of small drift velocity. These regimes are, with increasing time, the FDT-regime describing a local equilibrium within one of the valleys of the energy landscape, an intermediate plateau regime where the correlation function $q(t)$ stays close to the EA-order parameter $q_c$ and where the characteristic time scale $t_p(v)\sim v^\zeta$, and the asymptotic regime with characteristic time scale $t_a\sim v^{-1}$. In the creep phase $\cal A$ this asymptotic regime is ruled by two time scales, $t_a\sim v^{-1+\eta}$ and $t'_a\sim v^{-1-\eta}$. The exponent $\eta$ also determines the power law of $v(F)$. The numerical calculations have been performed for a wide range of velocities including values as small as $10^{-30}$ and over times ranging from $10^{-4}$ to $10^{36}$. This is necessary in order to deduce the full asymptotic behavior and even at these extreme values part of the structure is not yet fully developed. The second aspect relates to glassy non-eqilibrium dynamics of mean field models. The model investigated here has several advantages in this respect. It uses Lan\-ge\-vin dynamics which is certainly easier to handle than for instance Glauber dynamics in Ising type models. Depending on $\gamma$, continuous as well as discontinuous ergodicity breaking transitions are found. With applied external force a stationary non-equilibrium state is reached where correlation and response functions depend only on time differences. The inverse velocity $v^{-1}$ plays the role of an external long time scale. The replica treatment of this model \cite{Mez:manifold90,Mez:manifold91} predicts transitions between a phase with continuous replica symmetry breaking for $\gamma<\gamma_c=1$, a 1RSB-phase for $\gamma>1$ and $T<T_{c,\rm 1RSB}$, and a replica symmetric phase for $\gamma>1$ and $T>T_{c,\rm1RSB}$. The present phase diagram differs in the sense that $T_c>T_{c,\rm1RSB}$ and $\gamma_c>1$. A difference in $T_c$ has been observed before in models with discontinuous transitions \cite{KibachHo93I,KiTh87,ho:binperc92,CHS93}. This was traced back to the fact that the states contributing most to the static replica calculation are different from those relevant for dynamics and are not accessible within finite time in the thermodynamic limit. The same now appears to be true for continuous transitions as well. There have been several proposals regarding the long time dynamics. For the SK-model Sompolinsky and Zippelius \cite{Somp81,SZ82} proposed a hierarchical or ultrametric organization of long time scales. This can be rephrased as the postulate \cite{Ho:DySK84} that correlation and response functions can be expressed as functions of $x(t)=1-\ln t/\ln \bar t$ where $\bar t$ is some long external time scale. It has, however, been shown that this leads to inconsistencies \cite{Ho:DySK87}. In the present formulation this requirement means $\nu(t)=0$ for times where the hierarchy exists. This is not observed. The assumption of an ultrametric hierarchy of time scales leads to results which are identical to those obtained by Parisi's continuous replica symmetry breaking scheme \cite{Par79}. One of the quantities to be compared is the probability of overlaps $P(q)$ which is the derivative $P(q)=|\partial_q x(q)|$ of a function $x(q)$ which in dynamics is given by \begin{equation} x\big(q(t)\big)=\frac{1}{1+n(t)}\,. \end{equation} Within this scheme only $x(q)$ is determined, whereas correlation and response functions are not unique for long times. This is in contrast to the present investigation where correlation and response functions are unique for all times. The function $x(q)$ obtained from replica theory and from the present investigation are compared in Fig.\ref{x(q)}. The difference between the results shows again that different states are of relevance in dynamics and replica theory. \mfigure{xq-B} {Function $x(q)$ in phase $\cal B$, see text. The results obtained in the present investigation (solid) and from replica theory (dashed) are shown.}{x(q)} For systems with discontinuous transitions \cite{KibachHo93I,KiTh87,ho:binperc92,CHS93} the QFDT-solution has been proposed. It requires $n(t)\to n_Q$ for $t\sim \bar t$. This is actually found in phase $\cal A$. Within this scheme nothing can be said about the generation of internal long time scales. Furthermore, the resulting correlation and response functions are again not uniquely determined, contrary to the present results. There has been considerable interest on mean field dynamics of spin glasses and related systems with non-equilibrium initial conditions \cite{CuKu93,FrMe94a,FrMe94b,CuKu94,BCuKuPa95,CuDe95,CuDou95}. In the work of Cugliandolo and Dean \cite{CuDe95} an exact solution of the spherical SK-model is reported. It shows aging phenomena although a replica calculation of this model does not require replica symmetry breaking. The remaining papers deal with models which have phases with broken replica symmetry (1-step or continuous). Various proposals are made for the long time properties of correlation and response functions depending on two time arguments $t$ and $t'$. Again for $t-t'\to\infty$ and $t'\to\infty$ with finite $(t-t')/t'$ the resulting correlation and response functions could not be determined uniquely. This is likely to be due to an incomplete analysis of the plateau regime where $t-t'\sim {t'}^\zeta$ with $0<\zeta<1$. A comparison with the present results certainly has to be taken with care because the specific non-equili\-bri\-um situation is different. Nevertheless using $t'$ as external time scale one expects that at least the properties for times $t-t'\ll t'$ can be compared. This includes the plateau regime. A careful investigation of this regime and the kind of FDT-violation taking place there seems necessary in this case, too and this is likely to remove the arbitrariness in the correlation and response functions found so far.
\section*{Acknowledgments} For helpful discussions, we thank physicists at Cornell, DESY, Valencia, and at Montreux. This work was partially supported by U.S. Dept. of Energy Contract No. DE-FG 02- 96ER40291. \section*{Footnotes} \begin{enumerate} \item For testing for $(V+A)$ versus $(V-A)$, the use of $I_7$ for $\{\rho ^{-},\rho ^{+}\}$ gives less than a $1\%$ improvement over $I_4$ at $M_Z$, $10 GeV$, or $4 GeV$. If in addition the $\tau ^{-}$momentum direction is known via a SVX detector, there is only an $\sim 11\%$ improvement. The same numbers occur for $\{ {a_1}^{- },{a_1}^{+} \}$. In contrast, by using $I_4$, instead of the simpler 2 variable $I\left( E_\rho ,E_{\bar \rho }\right) $ spin- correlation function, there is about a factor of 8 improvement at $M_Z$. \item Note $\frac{m_b}{m_t}\sim \frac 5{174}\sim 3\%$, and $% \frac{m_\nu }{m_\tau }<\frac{23.8}{1777}\sim 1.4\%$ so this symmetry is badly broken in the masses for the 3rd family. However, for the other leptons this symmetry may be more strongly broken since $\frac{m_{\nu _e}}{m_e}<10^{-5}$, and $\frac{m _{\nu _\mu }}{m_\mu }<0.15\%$ from the current empirical bounds. From phenomenological mass formulas, e.g. \cite{har}, such as the GUT mass formula, $\nu_{\tau}$:$\nu_{\mu}$:$\nu_{e}$ $\sim$ ${m_t}^2$:${m_c}^2$:${m_u}^2$, the tau leptons are also the least asymmetric since then $\frac{m_{\nu _\tau}}{m_\tau} \approx 10^{-8}$, $\frac{m_{\nu _\mu}}{m_\mu} \approx 10^{-11}$, and $\frac{m_{\nu _e}}{m_e} \approx 3\cdot 10^{-14}$ for the normalization $m_{\nu_{\tau}}=20eV$. \item Details on the analysis of the $\tau \rightarrow \pi \nu$ modes will be reported elsewhere \cite{C94a}. \item The tests in this paper use ($\tau^{-} \tau^{+}$) spin- correlations as it is assumed that the $e^{-}$ and $e^{+}$ colliding beams are not longitudinally-polarized. Recently, Y.-S. Tsai \cite{ch1,ch2} has shown that in tau decays the sensitivities of tests for $CP$ violation, and for other types of ``new physics'', are substantially improved in regard to both systematic and statistical errors by the use of longitudinally- polarized beams at the ($\tau^{-} \tau^{+}$) threshold. \end{enumerate}
\section{Introduction} One of the outstanding questions in our understanding of fully-developed turbulence is the mechanisms by which the cascade of energy to small scales is maintained. That the cascade is intermittent is well recognized, but the phenomenological and dynamical models used to address the problem are rarely connected to the dynamics in the full Navier-Stokes equations. In this note it is shown that a popular model for explaining turbulent intermittency, the GOY model \cite{gledzer,yo}, shares some symmetries with terms in a decomposition of the spectral Navier-Stokes equations into the interactions between its helical components \cite{waleffe92}. The essential common property that is identified in both the GOY model and in Navier--Stokes is the importance of interactions between components with oppositely signed helicity. This new role for helicity supporting an intermittent cascade contrasts strongly with helicity's previously identified role in blocking the cascade \cite{andrelesieur77,polifkeshtilman89}, a role that was proposed earlier based on an analogy to the way magnetic helicity creates force-free states \cite{moffatt69}. In the GOY model with the standard parameters for three-dimensional turbulence, interactions between modes with oppositely signed helicity occur naturally as the sign of helicity reverses between neighboring shells \cite{leo_lohse_wang_benzi}. Whether Navier--Stokes turbulence follows a similar path is a more difficult question because there are several paths, characterized by interactions between different components of the helicity, that the cascade can follow. One way to investigate this question is to consider several variants of the GOY models that investigate each path in the full Navier--Stokes equations individually or in unison. One question would be how strongly the statistical behavior of the cascade depends on the symmetries in the different models. Another line of investigation is to determine which path the cascade follows in the full Navier--Stokes equations. In this note, preliminary results following both of these approaches are presented and the direction of a more complete study is presented. Also included will be new analysis of the Kerr-Siggia shell model \cite{kerrsiggia}, which has a cubic, non-positive definite invariant that shares some dynamical properties with helicity and served as an inspiration for part of this new proposal. In all of these cases it will be argued that competition between the transfer of energy and the transfer of the generalized helicity could explain the presence of numerically observed chaotic dynamics and intermittency in the energy cascade. For the GOY models, new results on intermittency corrections will be used to illustrate the importance of the second quadratic invariant in the energy cascade. In section 2, the new results for the Kerr-Siggia model will be reported. In section 3, the GOY model is reviewed and how the inviscid conserved quantities and the dynamics depend on the free parameters present in the model is discussed. In section 4, an argument that predicts the transition from a trivial dynamics (dominated by the presence of an attractive fixed point) to a fully chaotic regime for some critical values of the free parameters is discussed. Some numerical results for the energy transfer are also discussed. In section 5, new versions of the GOY model are introduced by considering explicitly the possibility of having shells which transport positive or negative helicity exactly as occurs in the Navier--Stokes equations. Two preliminary calculations with the full Navier--Stokes equations that support the importance of the interactions between components with oppositely-signed helicity are presented. Some problems that can be studied by using the new variant of the GOY model are discussed and some new analysis of the full Navier--Stokes equations that could be done to illuminate these properties is presented. \section{Pulse scaling of Kerr-Siggia} The Kerr-Siggia model\cite{kerrsiggia} is a shell model with one complex variable per shell, originating from a decimation of the possible interactions between triads in Burgers equation. With a simple forcing and eddy viscosity the equations were $$du_1/dt = \epsilon/u_1^* + 2iu_2u_1^*$$ \begin{equation} du_n/dt = k_{n-1}i(u_{n-1}^2 + 2u_{n+1}u_n^*) \label{eq:kseq} \end{equation} $$du_N/dt=k_{N-1}i(u_{N-1}^2 + \nu_e|u_N|u_N)$$ where $\epsilon$ was the average energy input and dissipation, $\nu_e$ was taken to be $2^{2/3}$ and $k_n=2^n$. Defining $$E_n={1\over 2}u_n u_n^* $$ \begin{equation} H_n=\Re(u_n^* u_{n-1}^2) \label{eq:ksconserve} \end{equation} $$A_n=\Im(u_n^* u_{n-1}^2)$$ for $\epsilon=\nu_e=0$ there are two inviscid invariants $E=\sum{E_n}$ and $H=\sum{H_n}$. The first is energy and the second, while not positive-definite, can be treated as a Hamiltonian with canonical variables $u_n$ and $u_{-n}=u^*_n$ as follows: \begin{equation} du_n/dt = ik_n\delta H/\delta u_{-n} \label{eq:kscanon} \end{equation} The energy transfer between shells is $\epsilon_n=-k_{n-1}A_n$. There is a trivial, unstable ``Kolmogorov'' solution of with $u_n=-i(2^{1/3}\epsilon)^{1/3}k_n^{-1/3}$, $H_n=0$, and $E_n=2^{2/9}\epsilon^{2/3}k_n^{-2/3}$ corresponding to a solution of an earlier cascade model\cite{des_novikov74} . The context for discussing this model along with the GOY model and Navier-Stokes is the extra invariant $H$. Despite the fact that this $H$ is cubic and not quadratic, and that neither Euler nor GOY has a Hamiltonian of this form, due to the non-positive definite nature of $H$, it appears to have some of the same qualitative effects upon the cascade that we speculate helicity is capable of for Navier-Stokes and GOY. The point is that while additional invariants can block the energy cascade, as enstrophy does in two dimensions and helicity does to some degree in Navier-Stokes\cite{andrelesieur77,polifkeshtilman89}, due to the non-positive definite nature of the invariant there is an escape route where the cascade can find a way around this blockage. As will be demonstrated for the Kerr-Siggia model, this can take the form of pulses. From the tools used to demonstrate this for the Kerr-Siggia model, it will then be demonstrated that there is weak evidence for analogous phenomena in Navier-Stokes. In the original discussion\cite{kerrsiggia}, two classes of solutions besides the trivial ``Kolmogorov'' solution were discussed. First, stationary solutions for a small number of shells with no forcing or dissipation ($\epsilon=\nu_e=0$) and maximal $H$ were discussed. Second, forced, dissipative solutions were discussed. Intermittency was found in the time dependent solutions and the effect of the extra invariant was noted, particularly as it affected the slope of the energy spectrum, which was $<E_n>\sim k_n^{-1/2}$ rather than the Kolmogorov solution, but what effect the stationary solutions might exert upon the forced, dissipative time-dependent solutions was not considered. For the present calculations, $\epsilon=1$ was chosen, which gives a characteristic timescale of $t=1$. Using as initial conditions $u_1=u_2=(1,1)$, $u_n=0$, $n\geq3$, $N=14$, it took until $t=3.9$ for the effects of initial transients to dissappear. Then statistics were taken until $t=6.8$. Figure 1 shows the spectra of $<E_n>$ and $<H_n>$ for this period as well as $k_n^{-1/2}$ curves, confirming the results of the original paper\cite{kerrsiggia}. Details will be discussed after the evidence for pulses is presented. Figure 2 shows $E_n$ and $H_n$ spectra for a series of moderately spaced times and the time development of $E$, $H$ and dissipation for this time period. By moderately spaced in time it is meant that the times shown are not so closely spaced so as to show continuous development, but are close enough to show a relationship between pulses of $E_n$ and $H_n$ and intermittent bursts in the dissipation. The primary event to focus upon is best illustrated in the $H_n$ spectra. In this sequence it starts as a pulse of positive $H_n$ centered on $n=4$ at $t=6.11$. It is associated with only one of several bumps in the energy spectrum at this time and is not associated with the spike in energy dissipation at $t=6.15$. This spike in energy dissipation is assocated with one of the higher shell bumps in the energy spectrum and comes from a pulse at an earlier time of oppositely signed $H_n$ similar to the pulse to be described. Following the appearance of the positive peak of $H_n$ in $n=4$ at $t=6.11$, this peak breaks off from the lower shells and slowly propagates to larger shell numbers. The energy peak associated with it moves in tandem. Spectra of the transfer rates of $E_n$ and $H_n$ have also peaks that move with the pulse. When the effects of the highest shell, where the dissipation occurs, are felt, the pulse stalls at $t=6.29$ before the energy in the pulse suddenly dissipates at $t=6.34$. The stalling is the probable source of the bump in the time averaged energy spectra just before the dissipation regime. While this bump is on top of a spectrum less steep than Kolmogorov ($k_n^{-1/2}$ rather than $k_n^{-2/3}$), it is qualitatively similar to a bump in the turbulent energy spectra for atmospheric observations\cite{champagneetal77}, spectral closures\cite{andrelesieur77} and forced calculations of Navier-Stokes turbulence\cite{kerr85}. For Navier-Stokes the bump is believed to be associated with a bottleneck effect\cite{LohseGroeling} where the decrease in the slope of kinetic energy spectrum in the dissipation regime blocks the free-flow of kinetic energy just at the boundary between the inertial and dissipation subranges. While this effect probably plays some role in the appearance of the bump in the Kerr-Siggia model, examination of figure 2 suggests a strong role for the stationary solutions associated with the second invariant. This comes from noticing that the cubic invariant is nearly maximal over the shells covered by the bump. While this pulse is dissipating at $t=6.34$, the next major pulse of negative $H_n$ is beginning to move into shell 2 and positive $H_n$ for the major pulse following that is developing in shell 0 from the forcing. So a succession of $E_n$ and alternately signed $H_n$ pulses is suggested. Clearly this is a simplified picture as there are minor spikes in dissipation between the major spikes that are associated with weak pulses with small $H_n$ of no particular sign. An example of such a weak pulse is the blip in $H_n$ at $n=8$ for $t=6.35$ and the rapidly moving $E_n$ at this time. To quantitatively demonstrate the alternation in sign of the strong pulses, figure 3 is a contour plot in shell and time separation of correlations between different shells and times of $E_n$ and $H_n$. These plots are similar to contour plots of the energy transfer in forced Navier-Stokes calculations\cite{kerr90} and also \cite{kida_when} and in meteorology are referred to as Hovm\"uller diagrams. These are: \begin{equation} <(F_{n+\Delta_n,t+\Delta t}- \overline{F_{n+\Delta_n}}) (F_{n,t}-\overline{F_{n}})> \label{eq:ksflux} \end{equation} where $F_n$ is either $E_n$ or $H_n$. Positive correlations are dark, negative are light. These plots are for $n=1$, the second shell. The effect of a single pulse is the first region of increasing $\Delta_n$ and $\Delta t$ originating at $(0,0)$. The propagation is linear after the first few shells. Starting at about $\Delta t=0.5$ there is another strong dark region in the $E_n$ correlation and a strong light region in the $H_n$ correlation. This supports the qualitative picture coming from watching the time development that there are a succession of pulses of oppositely signed $H_n$. The appearance of these pulses raises several questions. First, what modification of the stationary solutions can propagate as a unit? Second, what causes the alternation in sign of $H_n$ of the pulses, is it the forcing or is it the nonlinear dynamics? We will not attempt to answer these questions. The point we do want to make is that there is some connection between the alternation in sign of the extra non-positive definite conserved quantity that seems to be associated with the appearance of pulses in the energy cascade and with intermittency in the model. In calculations where the extra invariant is suppressed, intermittency dissappears. These ideas are supported by noting that the mechanism with which the conserved quantities are pumped in the system and removed from the system can influence the scaling laws in the inertial subrange\cite{Jon_Lee80,LevShe95}. An extreme example is a calculation of the Kerr-Siggia model with a Newtonian viscosity\cite{Jon_Lee80} where the extra conserved quantity is suppressed, there is a Kolmogorov spectrum and no intermittency. These are subtle questions that would require more accurate studies. How can the pulses be related to the spectra in figure 1? The $<E_n>$ spectrum in figure 1 goes as $2^{-n/2}$. By dimensional arguments one might expect that the $<H_n>$ spectrum would obey $2^{-3n/4}$, but this is not required since at any given time $<H_n>$ can have either sign. In fact, the $<H_n>$ spectrum is less steep than this and also seems to follow $2^{-n/2}$. To understand this, imagine that each pulse is a coherent package of $E_n$ and $H_n$ traversing the spectrum, spending on average $2^{-n/2}$ time in each shell. Then the time averaged spectra $<E_n>$ amd $<H_n>$ will both have $2^{-n/2}$ spectra. This is similar to the argument that has been used to generate a -5/3 spectrum from fluctuations in a strained Burgers vortex\cite{lundgren82}. If $\Delta t$ spent in each band goes as $2^{-n/2}$ as this suggests, then this would imply that the bands in the $(\Delta_n,\Delta t)$ plots should approach zero slope as $\Delta_n$ increases. There is some tendency in this direction for small $\Delta_n$ in figure 3, but for larger $\Delta_n$ when the stalling noted at $t=6.29$ in figure 2 and dissipation effects are important, the bands are nearly linear. Again, the time-averaged $<H_n>$ should not have any particular sign, as evidenced by shell 4 in figure 1, and their magnitude $|<H_n>|$ should decrease as the averaging time is lengthened. This has been verified by using different time intervals for the time averaging. \section{The GOY model} Given this discussion, let us now examine properties of the Kerr-Siggia model shared by the GOY model \cite{gledzer,yo}. The GOY model has a very rich dynamical behaviour and it has been the object of many studies in recent years (see \cite{leo_lohse_wang_benzi,jpv,bbp,pbcfv,bllp,gpz} for some numerical and analytical results). It is the most popular shell-model for 3 dimensional turbulence because of its intermittent properties are very close to the corresponding quantities in Navier-Stokes equations when the parameters of the nonlinear terms share some properties with the nonlinear term in the Navier-Stokes equations. In particular, for zero viscosity and no external forcing, when the system has the same conservation laws as a 3D flow: conservation of energy, of helicity and of volume in phase space. The dynamical equations are as follows: \begin{equation} \frac{d}{dt} u_n =i\, k_n \left(u^{*}_{n+1}u^{*}_{n+2} + b u^{*}_{n+1}u^{*}_{n-1} +c u^{*}_{n-1}u^{*}_{n-2} \right) -\nu k_n^2 u_n +\delta_{n,n_0}f \label{eq:shell} \end{equation} where $\nu$ is the viscosity and $f$ is a forcing acting on a large-scale-shell (for example, $n_0=1$) introduced to obtain a statistically stationary dynamical state. This model has interactions only between first and second-neighbor shells in the Fourier space. The two parameters $b,c$ in the nonlinear terms are chosen such as to conserve energy, $E= \sum_n |u_n|^2$, for any choice of $\lambda$. The most general choice of parameters is: \begin{equation} b= - \frac{\epsilon}{\lambda};\qquad c= -\frac{1-\epsilon}{\lambda^2} \label{eq:parameters} \end{equation} where $\epsilon$ is the second free parameter in the model. The GOY model also has a second quadratic invariant beside energy: \begin{equation} H = \sum_n \chi(\epsilon)^n k_n^{\alpha(\epsilon,\lambda)} |u_n|^2. \label{eq:helicity} \end{equation} While energy conservation is forced by the choice (\ref{eq:parameters}), the characteristics of the second invariant, $H$, change by changing the values of $\epsilon$ and $\lambda$. When $\epsilon < 1$ this second invariant is not positive-definite ($\chi(\epsilon) = -1$), while if $\epsilon > 1$ it is positive-definite ($\chi(\epsilon) = +1$). By remembering that the Navier-Stokes equations are characterized by having a second inviscid invariant that is positive-definite in 2D (enstrophy) and non-positive definite in 3D (helicity), the value $\epsilon=1$ can be identified as the border between a shell model for 2D turbulence ($\epsilon>1$) and a shell model for 3D turbulence ($\epsilon < 1$). In the following, the problem of whether shell models like GOY-model are a good representation of 2D turbulence\cite{angelo} is not addressed and only the range ($0<\epsilon<1$) where the dynamics should reproduce aspects of a 3D turbulent flow will be considered. By looking in detail at the structure of the second invariant, only when $\alpha(\epsilon,\lambda)=1$ does it have physical dimensions coinciding with Navier-Stokes helicity \cite{leo_lohse_wang_benzi}. This defines a line in the plane of free parameters where the inviscid conservation laws of the GOY model are very similar to the 3D Navier-Stokes equations. Because this invariant has the same physical dimensions as 3D helicity and it is non-positive, we will denote it as the {\it GOY-helicity} in the following. In the last section, a modified version of the GOY model will be introduced with a second invariant having more correspondance with fluid helicity. A necessary point before going on is that model (\ref{eq:shell}) has two inviscid fixed points corresponding to the Kolmogorov scaling $|u_n| \sim k_n^{-1/3}$ (constant flux of energy, zero flux of helicity) and to a fluxless scaling $|u_n| \sim k_n^{-(1+\alpha)/3}$ (constant flux of helicity, zero flux of energy) \cite{bllp}. For this study our interest in this model comes from the presence in the $(\epsilon,\lambda)$ plane of a region where the static Kolmogorov-like fixed point is dynamically unstable \cite{bllp}. The dynamics is fully chaotic and shows an intermittent cascade of energy toward small scales \cite{jpv,pbcfv} with a complex (multifractal) structure of the attractor in the phase-space. This intermittency is quantified by measuring the scaling exponents $\zeta(p)$ for the structure functions in the inertial range: \begin{equation} S_p(k_n) = <|u_n|^p> \sim k_n^{\zeta(p)} \end{equation} Only very recently \cite{leo_lohse_wang_benzi,gpz} has it been realized that the second quadratic invariant plays a crucial role in the dynamics of the model. In \cite{leo_lohse_wang_benzi}, it was found by varying the two free parameters $(\epsilon, \lambda)$ in the 3D-physically relevant region ($0< \epsilon <1; \lambda > 1$) that along the line of constant helicity ($\alpha(\epsilon,\lambda)=1$), the model has the same intermittent behaviour. That is, the set of $\zeta(p)$ depends only on the value of $\alpha$, giving, for the first time, numerical evidence that the dynamics of the model is strongly dependent on the presence of the second inviscid-invariant. Furthermore, it has been shown\cite{gpz} that by modifying the nonlinear term such as to destroy the presence of the second invariant (but still preserving the inviscid energy conservation) the intermittent corrections to K41 seem to weaken. \section{The transition to chaos} It has been shown\cite{bllp} that by fixing $\lambda$ and varying the $\epsilon$ parameters (and therefore, by changing $\alpha$) the GOY model undergoes a transition to chaos following a ``Ruelle-Takens'' scenario. In particular, there exists a critical value, $\epsilon_c$ such that for $\epsilon < \epsilon_c$ the Kolmogorov fixed point $u_n \sim k_n^{-1/3}$ is dynamically stable. Extending this analysis by changing the ratio between shells in the range $1< \lambda <3$ it is found that the ``Ruelle-Takens'' transition is quite general\cite{skl}: there exists a line in the plane ($\epsilon,\lambda$) which divides the region where a Kolmogorov-like fixed point is dynamically stable from a region where the dynamics is chaotic and intermittent. This qualitative trend of the transition line appears to beto be universal, even if the exact location can be slightly influenced by the forcing and by the value of viscosity. In the following, a very simple argument is presented based on the presence of the second non-positive invariant that predicts, with good accuracy, the existence of the transition and its location for any value in the plane ($\epsilon,\lambda$). Consider the two inviscid quadratic invariants, the energy and the generalized-helicity, and their currents. First, for zero viscosity and zero forcing energy conservation gives: \begin{equation} \frac{d}{dt}|u_n|^2 = J_{n-1} -J_{n}, \label{eq:flux_en} \end{equation} where the energy current $J_n$ \begin{equation} J_n = \Im[-\Delta_{n+1} -(1-\epsilon) \Delta_n]. \label{eq:current_en} \end{equation} is defined in terms of triple correlations: \begin{equation} \Delta_n = k_{n-1}u_{n-1}u_{n}u_{n+1}. \end{equation} The second conservation law for helicity takes the form: \begin{equation} \frac{d}{dt}(-)^{n}k_n|u_n|^2 = L_{n-1}- L_{n} \label{eq:flux_he} \end{equation} where the current of helicity, $L_n$, from the $n$th shell to the $n+1$th shell is: \begin{equation} L_n = (-)^{n} k_n \Im[\Delta_{n} -\Delta_{n+1}]. \label{eq:current_he} \end{equation} Let us suppose, for the moment, that there exists only one conserved quantity: energy. Then, very standard arguments \cite{orszag} tell us that if viscosity is zero, the system tends to equipartition, corresponding in the GOY model to $|u_n|^2 = const.$. If one switches on viscous effects, and starts with an initial configuration with energy concentrated in the first shells, an energy cascade toward small scales develops. This energy cascade has been interpreted as the attempt of the system to reach new equipartition state\cite{orszag}. This attempt at restoring equipartition is frustrated by viscous dissipation at small scales that continuously removes energy and prevents the small scales from reaching an equilibrium or quasi-equilibrium state. The Kolmogorov 1941 cascade, describing a smooth and constant transfer of energy from large scales to small scales is another way to rephrase this mechanism. But why does the flow not follow this picture of relaxing to a smooth and homogeneous transfer of energy and instead prefers to use a highly intermittent cascades consisting of bursts and blockages which are the origins of the intermittent corrections to the $\zeta(p)$ exponents? This is where we are proposing that the second inviscid quadratic quantity enters into the picture. It is well known that in 2D turbulence the presence of a second positive-definite quadratic invariant (enstrophy) does not allow the energy to cascade forward (toward small scales) \cite{kraichnan}. This is a general result; {\it either:} the two conserved quantities must transfer in different directions in Fourier space {\it or: } only one can transfer to small scales. For example, in 2D turbulence it is widely believed that there exists a forward transfer of enstrophy and a backward transfer of energy (inverse cascade). In contrast, the presence of a second non-positive definite quadratic invariant, like helicity in 3D turbulence, is only a minor constraint on the forward transfer of energy. Moreover, it is a constraint that, due to the non-positiveness, can have strong spatial and temporal fluctuations. If this picture is correct, intermittency in the 3D energy transfer could be the result of a competition between energy and helicity cascades. Temporal and/or spatial intermittency in the energy flux would be the result of switching between a net transfer of energy (possible due to cancellation effects in the helicity flux) and a depletion in the energy-transfer due to the presence of a non-zero helicity flux. How can these phenomenological ideas be checked in the GOY model? In the GOY model, a smooth and non-intermittent energy-transfer would correspond to dynamics near the Kolmogorov manifold $u_n \sim k_n^{-1/3}$. This implies that the energy flux (\ref{eq:flux_en}) is almost constant in the inertial range and that the helicity flux (\ref{eq:flux_he}) is almost vanishing. This Kolmogorov behaviour is obtained when the model has a static stable fixed point. It is natural, then, to ask if it is possible to understand the transition from the static behaviour to chaotic dynamics by invoking the second invariant. By plugging the Kolmogorov solutions into the expression for the generalized-helicity (\ref{eq:helicity}), we obtain: \begin{equation} H = \sum_n \chi(\epsilon)^n k_n^{\alpha(\epsilon,\lambda)-2/3}. \label{eq:helicity2} \end{equation} It is therefore clear that, whether the exponent $(\alpha-{2\over3})$ in (\ref{eq:helicity2}) is positive or negative determines whether $H$ receives most of its important contribution from small or large scales, respectively. Therefore, when $\alpha > 2/3$ the second invariant is concentrated at small scales and, as in 2D turbulence, prevents a smooth forward transfer of energy. This is reflected by strong intermittency and large deviations from Kolmogorov scaling. But, one can imagine that from time to time that it is still possible to transfer energy if some cancellation effects lead to an almost zero $H$-flux. On the other hand, when $\alpha < 2/3$ energy transfers toward small scales without having any relevant change in $H$, i.e. the model relaxes in to a trivial Kolmogorov like fixed point. Figure 4 shows the numerical results\cite{skl} for the transition from a static Kolmogorov behaviour to chaotic dynamics by changing $\epsilon$ and $\lambda$. As predicted, the transition happens near the critical line defined by: \begin{equation} \alpha(\epsilon_c,\lambda_c) = \frac{2}{3}, \qquad \lambda_c= (1-\epsilon_c)^{-3/2} \label{eq:prediction} \end{equation} The systematic shift of $5\%$ between the prediction (\ref{eq:prediction}) and the numerical results is probably due to viscosity as previously discussed\cite{skl} We believe that this very simple result is a new important confirmation that the dynamics of the model is strongly influenced by the second invariant. Figure 5 shows decaying numerical experiments, that is zero forcing and non-zero viscosity with energy initially concentrated at large scales, that compare the dynamical behaviour of the model in two characteristic regions: case A [$\epsilon=0.1, \lambda =2$], where there is smooth energy-transfer regime (on the right side of the critical curve in fig. 4) and case B [$\epsilon=0.5, \lambda =2$], where the dynamics is chaotic and intermittent (left side of the curve in fig. 4). What is interesting is that for case A, when the second invariant does not introduce any constraint in the energy transfer ($\alpha < 2/3$), the energy dissipation is a smooth function. This means that energy is transferred through the inertial subrange without any blocking and with a power-law behaviour in time. But, in case B ($\alpha > 2/3$), the energy dissipation has a staircase shape in time indicating long periods when little energy reaches small scales (blocking) interrupted by short bursts of dissipation. This would be consistent with the suggested role for helicity where in addition to blocking the transfer, cancellation in the helicity transfer is associated with strong dissipation events. Figure 6 shows the time evolution of the energy current $J_{n}$ and the helicity current $L_{n}$ through a shell in the inertial range where the model is chaotic. Obviously, the two fluxes are correlated, but the interesting fact is that when there is a burst of energy, the helicity flux has a sinusoidal shape, i.e. a net forward transfer of energy is only possible if the net averaged transfer of helicity is zero. Expanding upon the earlier proposal, this suggests blocking of energy transfer due to competition with helicity, interrupted by strong dissipation events made possible by brief, intermittent periods of large helicity fluctuations, but no net helicity flux. The dissipation events might then be associated with a strong dynamical coupling between modes with oppositely signed helicity, which would permit large helicity fluctuations, but no net helicity flux. \section{A new shell model} As described in the previous section, the structure of what we call helicity in the GOY model is only partially consistent with the helicity in the Navier-Stokes equations . Apart from the observation that it has the right dimensions and that it is not positive-definite, there is an asymmetry between odd and even shells that does not any counterpart in physical flows. One means of overcoming this problem is to introduce two dynamical variables in each shell, one transporting positive helicity $u_n^{+}$ and the other transporting negative helicity $u_n^{-}$. The next step is choosing how to couple these terms. For this, we will use a complete decomposition of the three dimensional Navier-Stokes eqs.\cite{waleffe92} into a basis where the two independent components of the velocity field at each wavenumber correspond to two pure helical waves. In such a basis there are 4 possible independent classes of triads interactions distinguished by the combination of helicity transported from each one of the three interacting modes. Let us fix, for simplicity, the three modes ${\bf q},{\bf p},{\bf k}$ such that $|{\bf k}| < |{\bf p}| < |{\bf q}|$, and call $u_{s_k}({\bf k}), u_{s_p}({\bf p}), u_{s_q}({\bf q})$ the three interacting modes, where $(s_k, s_p, s_q) = (\pm 1,\pm 1,\pm 1)$ refer to the sign of helicity in each mode. Then, it is simple to show that each triad can fall into one of the four following classes: \begin{enumerate} \item $(s_k, s_p, s_q)= (+,+,+)$, or $(-,-,-)$ \item $(s_k, s_p, s_q)= (+,+,-)$, or $(-,-,+)$ \item $(s_k, s_p, s_q)= (+,-,-)$, or $(-,+,+)$ \item $(s_k, s_p, s_q)= (-,+,-)$, or $(+,-,+)$ \end{enumerate} Following this decomposition one is led naturally to introduce 4 classes of shell models, each one corresponding to one of the four independent classes of triad interaction present in Navier-Stokes eqs. What is quite remarkable is that the original GOY model belongs to one of this classes (the fourth). To demonstrate this, we write the general equation for this class using positive helicity shells $u_n^+$ and negative helicity shells $u_n^-$. $$\frac{d}{dt} u^{+}_n =i\, k_n \left(u^{-}_{n+1}u^{+}_{n+2} + b u^{-}_{n+1}u^{-}_{n-1} +c u^{-}_{n-1}u^{+}_{n-2} \right)^* -\nu k_n^2 u^{+}_n +\delta_{n,n_0}f^{+} $$ \begin{equation} \frac{d}{dt} u^{-}_n =i\, k_n \left(u^{+}_{n+1}u^{-}_{n+2} + b u^{+}_{n+1}u^{+}_{n-1} +c u^{+}_{n-1}u^{-}_{n-2} \right)^* -\nu k_n^2 u^{-}_n +\delta_{n,n_0}f^{-}. \label{eq:pmshell} \end{equation} for which the conserved energy and helicity are given by: \begin{equation} E = \sum_n |u_n^{+}|^2 + |u_n^{-}|^2 \end{equation} \begin{equation} H = \sum_n k_n(|u_n^{+}|^2 - |u_n^{-}|^2) \end{equation} exactly as in the Fourier-helicity decomposition of Navier-Stokes equation \cite{waleffe92}. By noticing that in the original GOY model shells $n$ and $n+2$ have the same GOY-helicity, it can be seen that (\ref{eq:pmshell}) is formed from two masked and uncorrelated versions of the original GOY model for the dynamical evolution of the variables $(u_1^+,u_2^-,u_3^+,...,u_{2n-1}^+,u_{2n}^-)$ and $(u_1^-,u_2^+,u_3^-,...,u_{2n-1}^-,u_{2n}^+)$. Therefore (\ref{eq:pmshell}) has, by definition, the same behaviour as the previous model. {}From this, we think it is of primary importance to study in details the dynamical behaviour of the other three shell models (corresponding to the classes 1,2,3), allowing helicity to have all the dynamical interactions found in Navier-Stokes. Work in this direction is in progress and will be reported elsewhere \cite{bbkt}. \section{Navier-Stokes helicity} Up to this point only ideal shell models with extra non-positive definite invariants have been considered and how they might be extended to more closely resemble the helicity interactions in the Navier-Stokes equations. We would like to relate these ideas directly to the Navier-Stokes equations. As noted, earlier attempts at understanding the effects of helicity have emphasized its power to block the cascade\cite{andrelesieur77,polifkeshtilman89}. But, despite the blocking power of the extra invariant, the shell model calculations are indicating that the cascade can proceed through interactions between shells where the sign of the extra invariant is opposite. In a full calculation, we would also want to see what the effects of helicity in physical space are. For example, there are low Reynolds number Navier-Stokes calculations of how vortex rings link and unlink and can generate and destroy helicity\cite{ArefZad,MelanderHussain}. Because spectral properties were not analyzed and because of the low Reynolds numbers of these simulations, strong conclusions about the effects of helicity cannot be made from these calculations. But it can be said that even though the initial conditions contained large-scale helicity, small scale structures appeared and production of helicity from viscous effects was not strongly blocked. Therefore, initial conditions that contain significant large-scale helicity, but show more clearly how a cascade is not blocked, would be desireable. Whether or not production of small-scale helicity by anisotropies plays a role, as has been suggested\cite{FrischSheSulem_AKA87}, will not be our objective. Simulations of isotropic, homogeneous turbulence in a periodic box are traditionally initialized with a given spectrum, but the phases of individual wavenumbers is completely random. To test the effects of helicity we propose constraining the phases of the velocity components such that the helicity of the Fourier modes is not all of one sign, unlike the investigations of the blocking power of helicity\cite{polifkeshtilman89}. Since shell models are indicating strong effects from extra invariant fluctuations even when the average value is zero, this suggests initializing a Navier-Stokes calculation with net zero helicity. Several tests of this type have been done, all with one qualitatively similar feature. Helicity when it first appears in the spectra pops up in two neighboring bands of opposite sign, then the bands separate. Other than this qualitatively feature, the number of tests is not yet sufficiently large to make definitive statements. Two cases are shown in figure 7. Both cases are $64^3$ simulations where only a small number of modes in wavenumber band 4 were initialized. Each has the common feature noted, but a rich variety of additional features as well. In case A, each mode is initialized with maximal helicity, but otherwise the phases were chosen randomly (by hand) and the net helicity was zero. Very quickly the helicity picture changes. From a zero helicity spectrum, soon helicity of opposite sign appears in shells on opposite sides of the initial energy shell. For a short period, time sequences show that the helicity peak at higher wavenumber moves to small scales until it dissipates, leaving net helicity of the sign of the large scale peak. It is during this phase that relative dissipation rate (dissipation/energy) is largest. Therefore, through dissipation of helicity at small scales, large-scale helicity is generated. Once only the large-scale helicity peak is left, the blocking action of the helicity at large-scales becomes important and the dissipation rate is suppressed. In case D, the helicity of each mode is chosen to be zero by using free-slip boundary conditions in along central planes in the box. All the modes except one use the same free-slip symmetry plane, with the exception imposed to break this symmetry. From these initial conditions helicity does not initial grow around the initial $k=4$ wavenumber band, but around the resonance band at $k=8$. Note that once again helicity first appears in neighboring bands with opposite sign. Then time sequences show that the bands move towards opposite ends of the spectrum. Once again, dissipation is largest when the high wavenumber helicity band moves into the dissipation band and is annihilated, then decreases when large-scale helicity of only one sign remains. To make this more quantitative, more calculations need to be done and there needs to be analysis of helicity spectra and transfer properties. But what of a relationship to the shell models? Investigations of shell models with many more than 2 variables per shell seem to invariably lead to reductions in intermittency. Fully developed turbulence when viewed as a shell model has an infinite number of degrees of freedom and should in this sense not be intermittent. But it is, and this is understood as being due to coherent structures in physical space, which in Fourier space implies strong phase correlations. The phase correlations therefore prune the number of paths the cascade can take, returning us to simple models with few paths and strong intermittency, such as those presented here. Therefore, to get meaningful comparisons between 3D direct calculations and shell models, there must be some means of identifying the paths the cascade will follow and calculating statistics along these paths. Given the difficulty of attaining this, let us make some other suggestions. First, there needs to be further work on bi-dimensional correlations of wavenumber and time, similar to figure 3 here and in earlier analysis\cite{kerr90,kida_when}. The question with direct calculations is what quantities to use. It has been found\cite{kerr90} that energy transfer spectra have a strong signature. Helicity spectra and helicity transfer spectra need to be analyzed in the same manner. For simulations with a small number of initial modes, such as the examples just given, at least for short times statistics for modes formed by the initial interactions and their daughters could be studied. It is our hope that analysis of new Navier-Stokes simulations of this type, coupled with any new understanding fo the role of helicity coming from shell models, will provide new insight into the nature of the intermittent cascade of energy to small scales in turbulent flows. \vskip 0.3cm {\bf Acknowledgements} We thank Detlef Lohse, Leo Kadanoff and Norbert Schoerghofer for having communicated to us their numerical data before publication. We thank, also, Roberto Benzi, Uriel Frisch Detlef Lohse, Leo Kadanoff, Giovanni Paladin, Norbert Schoerghofer, Elisabetta Trovatore and Angelo Vulpiani for useful discussion. One of us (LB) has been partially supported by the EEC Contract ERBCHBICT941034. NCAR is supported by the National Science Foundation.
\section{Wave function of the Universe} Investigation of the wave function of the Universe goes back to the fundamental papers by Wheeler and DeWitt \cite{DeWitt}. However, for a long time it seemed almost meaningless to apply the notion of the wave function to the Universe itself, since the Universe is not a microscopic object. Only with the development of inflationary cosmology \cite{b14}--\cite{MyBook} it became clear that the whole Universe could appear from a tiny part of space of the Planck length (at least in the chaotic inflation scenario \cite{Chaotic}). Such a tiny piece of space can appear as a result of quantum fluctuations of metric, which should be studied in the context of quantum cosmology. Unfortunately, quantum cosmology is not a well developed science. This theory is based on the Wheeler-DeWitt equation, which is the Schr\"{o}dinger equation for the wave function of the Universe. However, Bryce DeWitt, one of the authors of this equation, in some of his talks emphasized that he is not particularly fond of it. This equation has many solutions, and at the present time the best method to specify preferable solutions of this equation (i.e. its boundary conditions) is based on the Euclidean approach to quantum gravity. This method is very powerful, but some of its applications are not well justified. In such cases this method may give incorrect answers, but rather paradoxically sometimes these answers appear to be correct in application to some other questions. Therefore it becomes necessary not only to solve the problem in the Euclidean approach, but also to check, using one's best judgement, whether the solution is related to the original problem or to something else. An alternative approach is based on the use of stochastic methods in inflationary cosmology. These methods allows one to understand such effects as creation of inflationary density perturbations, the theory of tunneling, and even the theory of self-reproduction of inflationary universe. Both Euclidean approach and stochastic approach to inflation have their limitations. However, despite all its problems, quantum cosmology is a very exciting science which changed dramatically our point of view on the structure and evolution of the Universe. We will begin our discussion with the issue of the Universe creation. According to classical cosmology, the Universe appeared from the singularity in a state of infinite density. Of course, when the density was greater than the Planck density $M_{\rm P}^4$ one could not trust classical Einstein equations, but in many cases there is no demonstrated need to study the Universe creation using the methods of quantum theory. For example, in the simplest versions of chaotic inflation scenario \cite{Chaotic} inflation, at the classical level, could begin directly in the initial singularity. However, in certain models, such as the Starobinsky model \cite{b14} or the new inflationary universe scenario \cite{New}, inflation cannot start in a state of infinite density. In such cases one may speculate about the possibility that inflationary universe appears due to quantum tunneling ``from nothing.'' The first idea how one can describe creation of inflationary universe ``from nothing'' was suggested in 1981 by Zeldovich \cite{Zeld} in application to the Starobinsky model \cite{b14}. His idea was qualitatively correct, but he did not propose any quantitative description of this process. A very important step in this direction was made in 1982 by Vilenkin \cite{NothVil}. He suggested to calculate the Euclidean action on de Sitter space with the energy density $V(\phi)$, \begin{equation} S_E = \int d^4 x \sqrt{-g}\left(-{R\over{16\pi G}}+V(\phi)\right) = -{3M_{\rm P}^4\over 8 V(\phi)} \ . \label{Action} \end{equation} This action was interpreted as the action on the tunneling trajectory describing creation of the Universe with the scale factor $a = H^{-1} = \sqrt{3 M_{\rm P}^2\over 8\pi V}$ from the state with $a = 0$. This would imply that the probability of quantum creation of the Universe is given by \begin{equation} {\cal P} \propto \exp (-S_E) = \exp \left({3 \over 8 V(\phi)}\right). \label{Vil1} \end{equation} (In the first three sections of this review we use the system of units with the Planck mass $M_{\rm P} = 1$.) A year later this result received a strong support when Hartle and Hawking reproduced it by a different though closely related method \cite{HH}. They argued that the wave function of the ``ground state'' of the Universe with a scale factor $a$ filled with a scalar field $\phi$ in the semi-classical approximation is given by \begin{equation}\label{E31} \Psi_0(a,\phi)\sim \exp\left(-S_E(a,\phi)\right)\ . \end{equation} Here $S_E(a,\phi)$ is the Euclidean action corresponding to the Euclidean solutions of the Lagrange equation for $a(\tau)$ and $\phi(\tau)$ with the boundary conditions $a(0)=a, \phi(0)=\phi$. The reason for choosing this particular wave function was explained as follows. Let us consider the Green's function of a particle which moves from the point $(0,t')$ to the point ${\bf x},t$: \begin{equation}\label{E32} <{\bf x},0|0, t'> = \sum_n \Psi_n ({\bf x})\Psi_n(0) \exp\left(iE_{n}(t-t')\right) = \int d{\bf x}(t) \exp\left(iS({\bf x}(t))\right)\ , \end{equation} where $\Psi_n$ is a complete set of energy eigenstates corresponding to the energies $E_n\geq 0$. To obtain an expression for the ground-state wave function $\Psi_0({\bf x})$, one should make a rotation $t \rightarrow -i\tau$ and take the limit as $\tau \rightarrow -\infty$\@. In the summation (\ref{E32}) only the term $n=0$ with the lowest eigenvalue $E_0 = 0$ survives, and the integral transforms into $\int dx(\tau)\exp(-S_E(\tau))$. Hartle and Hawking have argued that the generalization of this result to the case of interest in the semiclassical approximation would yield (\ref{E31}). The gravitational action corresponding to one half of the Euclidean section $S_4$ of de Sitter space with $a(\tau) = H^{-1}(\phi)\cos H\tau$ ($0\leq\tau\leq H^{-1}$) is negative, \begin{equation}\label{E33} S_E(a, \phi) = - \frac{3\pi}{4} \int d\eta\Bigl[\Bigl(\frac{da}{d\eta}\Bigr)^2 - a^2 + \frac{8\pi V}{3}a^4\Bigr] = - \frac{3}{16 V(\phi)}\ . \end{equation} Here $\eta$ is the conformal time, $\eta = \int {d\tau\over a(\tau)}$. Therefore, according to \cite{HH}, \begin{equation}\label{E34} \Psi_0(a,\phi)\sim \exp{\Bigl(-S_E(a,\phi)\Bigr)} \sim \exp\left(\frac{3 }{16V(\phi)}\right) \ . \end{equation} By taking a square of this wave function one again obtains eq. (\ref{Vil1}). The corresponding expression has a very sharp maximum as $V(\phi) \rightarrow 0$\@. This suggests that the probability of finding the Universe in a state with a large field $\phi$ and having a long stage of inflation should be strongly suppressed. Some authors consider it as a strong argument against the Hartle-Hawking wave function. However, nothing in the `derivation' of the Hartle-Hawking wave function tells that it describes initial conditions for inflation. The point of view of the authors of this wave function was not quite clear. They have written that their wave function gives the amplitude for the Universe to appear from nothing. On the other hand, Hawking emphasized \cite{Hawk300} that ``instead of talking about the Universe being created one should just say: the Universe is.'' This seems to imply that the Hartle-Hawking wave function was not designed to describe initial conditions at the moment of the Universe creation. Indeed, eq. (\ref{Vil1}) from the very beginning did not seem to apply to the probability of the Universe creation. The total energy of matter in a closed de Sitter space with $a(t) = H^{-1}\cosh Ht$ is greater than its minimal volume $\sim H^{-3}$ multiplied by $V(\phi)$, which gives the total energy of the Universe $E {\ \lower-1.2pt\vbox{\hbox{\rlap{$>$}\lower5pt\vbox{\hbox{$\sim$}}}}\ } M_{\rm P}^3/\sqrt V$. Thus the minimal value of the total energy of matter contained in a closed de Sitter universe {\it grows} when $V$ decreases. For example, in order to create the Universe at the Planck density $V \sim M_{\rm P}^4$ one needs no more than the Planckian energy $M_{\rm P} \sim 10^{-5}$ g. For the Universe to appear at the GUT energy density $V \sim M_X^4$ one needs to create from nothing the Universe with the total energy of matter of the order of $M_{\rm Schwarzenegger} \sim 10^2$ kg, which is obviously much more difficult. Meanwhile, eq. (\ref{Vil1}) suggests that it should be much easier to create a huge Universe with small $V$ but enormously large total energy rather than a small Universe with large $V$. This seems very suspicious. There is one particular place where the derivation (or interpretation) of eq. (\ref{Vil1}) could go wrong. The effective Lagrangian of the scale factor $a$ in (\ref{E33}) has a wrong sign. The Lagrange equations do not know anything about the sign of the Lagrangian, so we may simply change the sign before studying the tunneling. Only after representing the theory in a conventional form can we consider tunneling of the scale factor. But this then gives us the probability of quantum creation of the Universe \begin{equation} {\cal P} \propto \exp (-|S_E|) = \exp \left(-{3 \over 8 V(\phi)}\right). \label{E366} \end{equation} This equation predicts that a typical initial value of the field $\phi$ is given by $V(\phi)\sim 1$ (if one does not speculate about the possibility that $V(\phi) \gg 1)$, which leads to a very long stage of inflation. Originally I obtained this result by the method described above. However, because of the ambiguity of the notion of tunneling from the state $a = 0$, I was not quite satisfied and decided to look at it from the perspective of derivation of the Hartle-Hawking wave function. In this approach the problem of the wrong sign of the Lagrangian appears again, though in a somewhat different form. Indeed, the total energy of a closed Universe is zero, being a sum of the positive energy of matter and the negative energy of the scale factor $a$.) Thus, the energy $E_n$ of the scale factor is negative, and in order to suppress terms with large negative $E_n$ and to obtain $\Psi_0$ from (\ref{E33}) one should rotate $t$ not to $-i\tau$, but to $+i\tau$. This gives \cite{Creation} \begin{equation}\label{E35a} \Psi_0(a,\phi) \sim \exp\Bigl(-|S_E(a,\phi)|\Bigr) \sim\exp \left(- \frac{3}{16V(\phi)}\right)\ , \end{equation} and \begin{equation}\label{E36} P(\phi) \sim|\Psi_0(a,\phi)|^2\sim \exp\Bigl(-2|S_E(a,\phi)|\Bigr) \sim\exp \left(- \frac{3}{8V(\phi)}\right) \ . \end{equation} Few months later this equation was also derived by Zeldovich and Starobinsky \cite{ZelStar}, Rubakov \cite{Rubakov}, and Vilenkin \cite{Vilenkin} using the methods similar to the first method mentioned above (investigation of tunneling in the theory with the wrong sign of the Lagrangian). The corresponding wave function (\ref{E35a}) was called ``the tunneling wave function.'' An obvious objection against this result is that it may be incorrect to use different ways of rotating $t$ for quantization of the scale factor and of the scalar field. However, the idea that a consistent quantization of matter coupled to gravity can be accomplished by a proper choice of a complex contour of integration may be too optimistic. We know, for example, that despite many attempts to suggest Euclidean formulation for nonequilibrium quantum statistics or for the field theory in a nonstationary background, such formulation does not exist. It is quite clear from (\ref{E32}) that the $t \rightarrow -i\tau$ trick would not work if the spectrum $E_n$ were not bounded from below. Absence of equilibrium, of any simple stationary ground state seems to be a typical situation in quantum cosmology. In some cases where a stationary or quasistationary ground state does exist, eq. (\ref{Vil1}) may be correct, see the next Section. In a more general situation it may be very difficult to obtain any simple expression for the wave function of the Universe. However, in certain limiting cases this problem is relatively simple. For example, at present the scale factor $a$ is very big and it changes very slowly, so one can consider it to be a C-number and quantize matter fields only by rotating $t \rightarrow -i\tau$\@. On the other hand, in the inflationary universe the evolution of the scalar field is very slow; during the typical time intervals $O(H^{-1})$ it behaves essentially as a classical field, so one can describe the process of the creation of an inflationary universe filled with a homogeneous scalar field by quantization of the scale factor $a$ only and by rotation $t \rightarrow i\tau$. Derivation of equations (\ref{Vil1}), (\ref{E36}) and their interpretation is far from being rigorous, and therefore even now it remains the subject of debate. In our opinion, the Hartle-Hawking wave function describes not the Universe creation, but the fluctuations of the Universe near its de Sitter ground state, under the condition that such a state exists, see next section. Meanwhile the distribution (\ref{E36}) is related to the probability of creation of inflationary universe from nothing (or from the space-time foam). However, the two different derivations of this probability distribution emphasize two slightly different features of the process. Investigation of tunneling should give the probability of quantum creation of the Universe of a size $H^{-1}$ from the Universe with $a = 0$. Meanwhile wave function of the ``ground state'' should give information about some kind of probability distribution of various Universes in the space-time foam. We will not concentrate here on this subtle difference since we believe that it would bring us far away from the domain of applicability of our approach. Also, we should emphasize again that quantum tunneling is necessary only if one cannot use the classical trajectory. In the Starobinsky model \cite{b14}, as well as in the new inflationary universe scenario \cite{New}, creation of the Universe ``from nothing'' appears to be one of the most natural mechanisms for inflation to begin. Meanwhile, in the simplest version of chaotic inflation scenario the process of inflation formally may begin directly in the singularity, in a state with infinitely large $V(\phi)$, without any need for quantum tunneling. However, quantum tunneling in that case is possible as well, since for $V(\phi) \sim 1$ the probability of quantum creation of inflationary universe is not exponentially suppressed. In the next section we will discuss stochastic approach to quantum cosmology. Within this approach equations (\ref{Vil1}) and (\ref{E36}) can be derived in a much more clear and rigorous way, but they will have somewhat different interpretation. \ \section{Wave function of the Universe and stochastic approach to inflation} The first models of inflation were based on the standard assumption of the big bang theory that the Universe was created at a single moment of time in a state with the Planck density, and that it was hot and large (much larger than the Planck scale $M_{\rm P}^{-1} =1$) from the very beginning. The success of inflation in solving internal problems of the big bang theory apparently removed the last doubts concerning the big bang cosmology. Even in our quantum mechanical treatment of the Universe production we still used the standard idea that the Universe as a whole can be described by one scale factor $a$, and its creation should be considered as a process beginning from $a = 0$. Meanwhile during the last ten years the inflationary theory has broken the umbilical cord connecting it with the old big bang theory, and acquired an independent life of its own. For the practical purposes of description of the observable part of our Universe one may still speak about the big bang. However, if one tries to understand the beginning of the Universe, or its end, or its global structure, then some of the notions of the big bang theory become inadequate. For example, in the chaotic inflation scenario \cite{Chaotic} even without taking into account quantum effects there was no need to assume that the whole Universe appeared from nothing at a single moment of time associated with the big bang, that the Universe was hot from the very beginning and that the inflaton scalar field $\phi$ which drives inflation originally occupied the minimum of its potential energy. On the other hand, it was found that if the Universe in this scenario contains at least one inflationary domain of a size of horizon (`$h$-region') with a sufficiently large and homogeneous scalar field $\phi$, then this domain will permanently produce new $h$-regions of a similar type due to quantum fluctuations \cite{b19,b20}. This process occurs in the old, new and extended inflation scenario as well \cite{b51}--\cite{EtExInf}. Thus, instead of a single big bang producing a one-bubble Universe, we are speaking now about inflationary bubbles producing new bubbles, producing new bubbles, {\it ad infinitum}. In this sense, inflation is not a short intermediate stage of duration $\sim 10^{-35}$ seconds, but a self-regenerating process, which occurs in some parts of the Universe even now, and which will continue without end. It is extremely complicated to describe an inhomogeneous self-reproducing Universe. Fortunately, there is a particular kind of stationarity of the process of the Universe self-reproduction which makes things more regular. Due to the no-hair theorem for de Sitter space, the process of production of new inflationary domains occurs independently of any processes outside the horizon. This process depends only on the values of the fields inside each $h$-region of radius $H^{-1}$. Each time a new inflationary $h$-region is created during the Universe expansion, the physical processes inside this region will depend only on the properties of the fields inside it, but not on the `cosmic time' at which it was created. In addition to this most profound stationarity, which we will call {\it local stationarity}, there may also exist some simple stationary probability distributions which may allow us to say, for example, what the probability is of finding a given field $\phi$ at a given point. To examine this possibility one should consider the probability distribution $P_c(\phi, \chi,t)$, which describes the probability of finding the field $\phi$ at a given point at a time $t$, under the condition that at the time $t=0$ the field $\phi$ at this point was equal to $\chi$. The same function also describes the probability that the scalar field which at time $t$ was equal to $\phi$, at some earlier time $t=0$ was equal to $\chi$. The probability distribution $P_c$ is in fact the probability distribution per unit volume in {\it comoving coordinates} (hence the index $c$ in $P_c$), which do not change during expansion of the Universe. By considering this probability distribution we neglect the main source of the self-reproduction of inflationary domains, which is the exponential growth of their volume. Therefore, in addition to $P_c$, we introduced the probability distribution $P_p(\phi,\chi,t)$, which describes the probability to find a given field configuration in a unit physical volume \cite{b19}. In the situations where one of these distributions can be stationary, we will speak about {\it global stationarity}. Let us remember some details of stochastic approach to inflation. Consider the simplest model of chaotic inflation based on the theory of a scalar field $\phi$ minimally coupled to gravity, with the effective potential $V(\phi)$. If the classical field $\phi$ is sufficiently homogeneous in some domain of the Universe, then its behavior inside this domain is governed by the equation $3H\dot\phi = -dV/d\phi$, where $H^2 = \frac{8\pi V(\phi)}{3 }$. Investigation of these equations has shown that in all power-law potentials $V(\phi)\sim \phi^n$ inflation occurs at $\phi > \phi_e \sim n/6$. In the theory with an exponential potential $V(\phi)\sim e^{\alpha \phi}$ inflation ends only if we bend down the potential at some point $\phi_e$; for definiteness we will take $\phi_e = 0$ in this theory. Inflation stretches all initial inhomogeneities. Therefore, if the evolution of the Universe were governed solely by classical equations of motion, we would end up with an extremely smooth Universe with no primordial fluctuations to initiate the growth of galaxies. Fortunately, new density perturbations are generated during inflation due to quantum effects. The wavelengths of all vacuum fluctuations of the scalar field $\phi$ grow exponentially in the expanding Universe. When the wavelength of any particular fluctuation becomes greater than $H^{-1}$, this fluctuation stops oscillating, and its amplitude freezes at some nonzero value $\delta\phi (x)$ because of the large friction term $3H\dot{\phi}$ in the equation of motion of the field $\phi$\@. The amplitude of this fluctuation then remains almost unchanged for a very long time, whereas its wavelength grows exponentially. Therefore, the appearance of such a frozen fluctuation is equivalent to the appearance of a classical field $\delta\phi (x)$ that does not vanish after averaging over macroscopic intervals of space and time. Because the vacuum contains fluctuations of all wavelengths, inflation leads to the creation of more and more perturbations of the classical field with wavelengths greater than $H^{-1}$\@. The average amplitude of such perturbations generated during a time interval $H^{-1}$ (in which the Universe expands by a factor of e) is given by \begin{equation}\label{E23} |\delta\phi(x)| \approx \frac{H}{2\pi}\ . \end{equation} The phases of each wave are random. Therefore, the sum of all waves at a given point fluctuates and experiences Brownian jumps in all directions in the space of fields. One can describe the stochastic behavior of the inflaton field using diffusion equations for the probability distribution $P_c(\phi,t|\chi)$. The first equation is called the backward Kolmogorov equation, \begin{equation} \label{Starb} \frac{\partial P_c(\phi,t|\chi)}{\partial t} = \frac{H^{3/2}(\chi)}{8\pi^2}\, \frac{\partial }{\partial\chi} \left({H^{3/2}(\chi)} \, \frac{\partial }{\partial\chi} P_c(\phi,t|\chi) \right) - \frac{V'(\chi)}{3H(\chi)} \frac{\partial }{\partial\chi} P_c(\phi,t|\chi) \ . \end{equation} In this equation one considers the value of the field $\phi$ at the time $t$ as a constant, and finds the time dependence of the probability that this value was reached during the time $t$ as a result of diffusion of the scalar field from different possible initial values $\chi \equiv \phi(0)$. The second equation is the adjoint to the first one; it is called the forward Kolmogorov equation, or the Fokker-Planck equation \cite{Star}, \begin{equation}\label{E3711} \frac{\partial P_c(\phi,t|\chi)}{\partial t} = \frac{\partial }{\partial\phi} \left( \frac{H^{3/2}(\phi)}{8\pi^2}\, \frac{\partial }{\partial\phi} \Bigl( {H^{3/2}(\phi)} P_c(\phi,t|\chi) \Bigr) + \frac{V'(\phi)}{3H(\phi)} \, P_c(\phi,t|\chi)\right) \ . \end{equation} One may try to find a stationary solution of equations (\ref{Starb}), (\ref{E3711}), assuming that $\frac{\partial P_c(\phi,t|\chi)}{\partial t} = 0$. The simplest stationary solution (subexponential factors being omitted) would be \cite{Star,Mijic,LLM} \begin{equation}\label{E38a} P_c(\phi,t|\chi) \sim \exp\left({3\over 8 V(\phi)}\right)\cdot \exp\left(-{3\over 8 V(\chi)}\right) \ . \end{equation} This looks like a miracle: The first term in this expression is equal to the square of the Hartle-Hawking wave function of the Universe (\ref{Vil1}), whereas the second one gives the square of the tunneling wave function (\ref{E36})! And we obtained this result without any ambiguous considerations based on Euclidean approach to quantum cosmology! At first glance, this result gives a direct confirmation and a simple physical interpretation of both the Hartle-Hawking wave function of the Universe {\it and} the tunneling wave function. First of all, we see that the distribution of the probability to find the Universe in a state with the field $\phi$ is proportional to $\exp\left({3\over 8 V(\phi)}\right)$. Note that we are speaking here about the state of the Universe rather than the probability of its creation. Meanwhile, the probability that the Universe emerged from the state with the field $\chi$ is proportional to $\exp\left(-{3\over 8 V(\chi)}\right)$. Now we are speaking about the probability that a given part of the Universe was created from the state with the field $\chi$, and the result coincides with our result for the probability of the quantum creation of the Universe, eq. (\ref{E36}). This would be a great peaceful resolution of the conflict between the two wave functions. However, the situation is much more complicated. In all realistic cosmological theories, in which $V(\phi)=0$ at its minimum, the Hartle-Hawking distribution $\exp\left({3\over 8 V(\phi)}\right)$ is not normalizable. The source of this difficulty can be easily understood: any stationary distribution may exist only due to compensation of the classical flow of the field $\phi$ downwards to the minimum of $V(\phi)$ by the diffusion motion upwards. However, diffusion of the field $\phi$ discussed above exists only during inflation. Thus, there is no diffusion motion upwards from the region $\phi < \phi_e$. Therefore expression (\ref{E38a}) is not a true solution of equation (\ref{E3711}); all solutions with the proper boundary conditions at $\phi = \phi_e$ (i.e. at the end of inflation) are non-stationary (decaying) \cite{b19}. It is possible to use the solution (\ref{E38a}) in the cases where the state can be quasistationary. For example, in the case when the effective potential has a local minimum with a sufficiently large $V$, this distribution gives a correct expression for the probability of the Hawking-Moss tunneling \cite{Star}. We were unable to find a situation in the context of inflationary cosmology where one could ascribe a more fundamental meaning to the Hartle-Hawking wave function, but of course this might be a result of our own limitations. One can get an additional insight by investigation of the probability distribution $P_p$. In order to do so, one should write stochastic equations for ${\cal V}(\phi,t|\chi)$, where ${\cal V}(\phi,t|\chi)$ is the total volume of domains with the field $\phi$ originated from the domains containing field $\chi$. The system of stochastic equations for ${\cal V}(\phi,t|\chi)$ can be obtained from eqs. (\ref{Starb}), (\ref{E3711}) by adding the term $3H{\cal V}$, which appears due to the growth of physical volume of the Universe by the factor $1 + 3H(\phi)\, dt$ during each time interval $dt$ \cite{Nambu}--\cite{LLM}: \begin{equation} \label{Starbx} \frac{\partial {\cal V}}{\partial t} = \frac{H^{3/2}(\chi)}{8\pi^2}\, \frac{\partial }{\partial\chi} \left( {H^{3/2}(\chi)} \, \frac{\partial {\cal V}}{\partial\chi} \right) - \frac{V'(\chi)}{3H(\chi)} \frac{\partial {\cal V}}{\partial\chi} +3H(\chi) {\cal V} \ , \end{equation} \begin{equation}\label{E372} \frac{\partial {\cal V}}{\partial t} = \frac{\partial }{\partial\phi} \left( \frac{H^{3/2}(\phi)}{8\pi^2}\, \frac{\partial }{\partial\phi} \Bigl( {H^{3/2}(\phi)} {\cal V} \Bigr) + \frac{V'(\phi)}{3H(\phi)} \, {\cal V}\right) + 3H(\phi) {\cal V}\ . \end{equation} To find solutions of these equations one must specify boundary conditions. Behavior of solutions of these equations typically is not very sensitive to the boundary conditions at the boundary $\phi = \phi_e$ where inflation ends; it is sufficient to assume that the diffusion coefficient (and, correspondingly, the first terms in the r.h.s. of equations (\ref{Starbx}), (\ref{E372})) vanish for $\phi < \phi_e$. The conditions at the Planck boundary $\phi = \phi_p$ play a more important role. In what follows we will assume that inflation ceases to exist at $\phi > \phi_p$ \cite{LLM}. This leads to the boundary condition \begin{equation}\label{PlanckBound} {\cal V}(\phi_p,t|\chi)= {\cal V}(\phi,t|\chi_p) = 0 \ , \end{equation} where $V(\phi_p) \equiv V(\chi_p) = O(1)$. One may try to obtain solutions of equations (\ref{Starbx}), (\ref{E372}) in the form of the following series of system of eigenfunctions of the pair of adjoint linear operators defined by the left hand sides of the equations below: \begin{equation} \label{eq14} {\cal V}(\phi,t|\chi) = \sum_{s=1}^{\infty} { e^{\lambda_s t}\, \psi_s(\chi)\, \pi_s(\phi) } \ . \end{equation} Indeed, this gives us a solution of eq. (\ref{E372}) if \begin{equation} \label{eq15} \frac{H^{3/2}}{8\pi^2} \frac{\partial }{\partial\chi} \left({H^{3/2}} \frac{\partial }{\partial\chi} \psi_s(\chi) \right) - \frac{V'}{3H} \frac{\partial }{\partial\chi} \psi_s(\chi) + 3H \cdot \psi_s(\chi) = \lambda_s \, \psi_s(\chi) \ , \end{equation} and \begin{equation} \label{eq17} \frac{\partial }{\partial\phi} \left( \frac{H^{3/2}}{8\pi^2} \frac{\partial }{\partial\phi} \left({H^{3/2}} \pi_j(\phi) \right) \right) + \frac{\partial }{\partial\phi} \left( \frac{V'}{3H} \, \pi_j(\phi) \right) + 3H\cdot \pi_j(\phi) = \lambda_j \, \pi_j(\phi) \ . \end{equation} In our case (with regular boundary conditions) one can easily show that the spectrum of $\lambda_j$ is discrete and bounded from above. Therefore the asymptotic solution for ${\cal V}(\phi,t|\chi)$ in the limit $t \rightarrow \infty$ is given by \begin{equation} \label{eq22} {\cal V}(\phi,t|\chi) = e^{\lambda_1 t}\, \psi_1(\chi) \, \pi_1(\phi)\, \cdot \left(1 + O\left( e^{-\left(\lambda_1 - \lambda_2 \right) t} \right) \right) \ . \end{equation} Here $\psi_1(\chi)$ is the only positive eigenfunction of eq. (\ref{eq15}), $\lambda_1$ is the corresponding (real) eigenvalue, and $\pi_1(\phi)$ is the eigenfunction of the conjugate operator (\ref{eq17}) with the same eigenvalue $\lambda_1$\@. Note, that $\lambda_1$ is the highest eigenvalue, $\mbox{Re} \left( \lambda_1 - \lambda_2 \right) > 0 $\@. This means that the distribution \begin{equation} \label{eq22aa} {P}_p(\phi,t|\chi) = e^{-\lambda_1 t} \,{\cal V}(\phi,t|\chi) \end{equation} gradually converges to the time-independent normalized distribution \begin{equation} \label{eq22a} {P}_p(\phi,\chi) \equiv {P}_p(\phi,t \rightarrow \infty|\chi) = \psi_1(\chi) \, \pi_1(\phi) \ {}. \end{equation} It is this stationary distribution that we were looking for. $ {P}_p(\phi,\chi)$ gives the fraction of the volume of the Universe occupied by the field $\phi$, under the condition that the corresponding part of the Universe at some time in the past contained the field $\chi$. The remaining problem is to find the functions $\psi_1(\chi)$ and $\pi_1(\phi)$, and to check that all assumptions about the boundary conditions which we made on the way to eq. (\ref{eq22}) are actually satisfied. We have solved this problem for chaotic inflation in a wide class of theories including the theories with polynomial and exponential effective potentials $V(\phi)$ and found the corresponding stationary distributions \cite{LLM}. Here we will present some of our results for the theory ${\lambda\over 4}\phi^4$. Solution of equations (\ref{eq15}) and (\ref{eq17}) for $\psi_1(\chi)$ and $\pi_1(\phi)$ shows that these functions are extremely small at $\phi\sim \phi_e$ and $\chi\sim \chi_e$, where $\phi_e\sim \chi_e \sim 1$ correspond to the end of inflation. These functions grow at large $\phi$ and $\chi$, then rapidly decrease, and vanish at $\phi = \chi= \phi_p$. With a decrease of $\lambda$ the solutions become more and more sharply peaked near the Planck boundary. To give a physical interpretation to our solutions, it will be convenient to parametrize $\lambda_1$ in the following form: $ \lambda_1= d(\lambda) H_{\rm \max}(\lambda)$. Here $d$ is the so-called fractal dimension of inflationary universe \cite{ArVil,LLM}, and $H_{\rm \max} $ is the maximal value of the Hubble constant in the model under consideration. For example, in the models where inflation ceases to exist at the Planck density $V(\phi) = 1$ the maximal value of the Hubble constant is given by $2\sqrt{2\pi\over 3}$. The eigenvalues $d(\lambda)$ corresponding to different coupling constants $\lambda$ are given by the following table: \begin{center} \begin{tabular}{|c|c|c|c|c|c|c|c|} \hline \hline $\lambda$ & $1$ & $10^{-1}$ & $10^{-2}$ & $10^{-3}$ & $10^{-4}$ & $10^{-5}$ & $10^{-6}$ \\ \hline $d$ & 0.9719 & 1.526 & 1.915 & 2.213 & 2.438 & 2.604 & 2.724 \\ \hline \hline \end{tabular} \end{center} As we see, in the limit $\lambda \to 0$ the fractal dimension $d(\lambda)$ grows toward the usual space dimension $3$. It is very interesting to study the behavior of $P_p$ at small $\phi$ and $\chi$, i.e. at the stage which determines the structure of the observable part of the Universe. One could expect to find a dependence similar to the one given by eq. (\ref{E38a}), i.e. $P_p\sim \exp\left({3\over 8 V(\phi)}\right)\cdot \exp\left(-{3\over 8 V(\chi)}\right)$. Indeed, this remains true for the dependence of $P_p$ on $\chi$. Meanwhile, since there is no diffusion term at $\phi < \phi_e$, the solution at small $\phi> \phi_e$ should match the solution obtained by neglecting the first (diffusion) term at $\phi < \phi_e$. As a result, instead of the product of the Hartle-Hawking and the tunneling solution for the theory ${\lambda\over 4} \phi^4$ for small $\phi$ and $\chi$ (for $\phi, \chi < \lambda^{-1/8}$) we obtain \begin{equation}\label{SMALLPHI} {\cal V}(\phi,\chi,t)\, =\, e^{ d(\lambda) H_{\rm \max}t}~ {P}_p(\phi,\chi)\, \sim\, e^{ d(\lambda) H_{\rm \max}t}~ \phi^{\sqrt{6\pi\over \lambda}\lambda_1}~ \exp\left(-{3\over 8 V(\chi)}\right) \ . \end{equation} Thus, the square of the tunneling wave function is here, but the square of the Hartle-Hawking wave function dropped away. The dependence of ${\cal V}(\phi,\chi,t)$ on $\chi$ and $\phi$ is extremely sharp. For example, for the realistic value $\lambda \sim 10^{-13}$ one has $ {P}_p(\phi,\chi) \sim e^{10^{13}\chi^{-4}} \, \phi^{10^8}$. The factor $e^{ d(\lambda) H_{\rm \max}t}$ controls the speed of exponential expansion of the volume filled by a given field $\phi$. {\it This speed does not depend on the field $\phi$}, and has the same order of magnitude as the speed of expansion at the Planck density. One should emphasize that the factor $e^{ d(\lambda) H_{\rm \max}t}$ gives the rate of growth of the combined volume of all domains with a given field $\phi$ (or of all domains containing matter with a given density) {\it not only at very large $\phi$, where quantum fluctuations are large, but at small $\phi$ as well, and even after inflation} \cite{LLM}. This result may seem absolutely unexpected, since the volume of each particular inflationary domain grows like $e^{3H(\phi)t}$, and after inflation the law of expansion becomes completely different. One should distinguish, however, between the growth of each particular domain, accompanied by a decrease of density inside it, and the growth of the total volume of all domains containing matter with a given (constant) density. In the standard big bang theory the second possibility did not exist, since the energy density was assumed to be the same in all parts of the Universe (``cosmological principle''), and it was not constant in time. The reason why there is a universal expansion rate $e^{\lambda_1 t}$ can be understood as follows. Because of the self-reproduction of the Universe there always exist many domains with $\phi \sim \phi_{p}$, and their combined volume grows almost as fast as $e^{3H(\phi_p) t}$. Then the field $\phi$ inside some of these domains decreases. The total volume of domains containing some small field $\phi$ grows not only due to expansion $\sim e^{3H(\phi_p) t}$, but mainly due to the unceasing process of expansion of domains with large $\phi$ and their subsequent rolling (or diffusion) towards small $\phi$. The distribution ${P}_p(\phi, \chi) = \psi_1(\chi) \, \pi_1(\phi)$ which we have obtained does not depend on time $t$. However, in general relativity one may use many different time parametrizations, and the same physics can be described differently in different `times'. One of the most natural choices of time in the context of stochastic approach to inflation is the time $\tau = \ln{{a\left(x, t \right) \over a(x,0)}} = \int{H(\phi(x,t),t)\ dt}$ \cite{Star,Bond}. Here $a\left(x, t \right)$ is a local value of the scale factor in the inflationary universe. By using this time variable, we were able to obtain not only numerical solutions to the stochastic equations, but also simple asymptotic expressions describing these solutions. For example, for the theory ${\lambda\over 4 } \phi^4$ both the eigenvalue $\lambda_1$ and the `fractal dimension' $d_f$ (which in this case refers both to the Planck boundary at $\phi_p$ and to the end of inflation at $\phi_e$) are given by $d_f = \lambda_1 \sim 3-1.1\, \sqrt \lambda$, and the stationary distribution is \cite{LLM} \begin{eqnarray} {P}_p(\phi,\chi) &\sim & \exp\Bigl(-{3\over 8V(\chi)}\Bigr)\, \Bigl({1\over V(\chi)+0.4} - {1\over1.4}\Bigr)\, \cdot \, \phi \,\exp\Bigl(-{\pi\, (3-\lambda_1)\phi^2}\Bigr) \nonumber \\ &\sim & \exp\Bigl(-{3\over 2 \lambda\, \chi^4}\Bigr)\, \Bigl({4\over \lambda \chi^4+1.6} - {1\over1.4}\Bigr)\, \cdot \, \phi \, \exp\Bigl(- 3.5 \sqrt\lambda\phi^2\Bigr)\ . \end{eqnarray} The first factor again coincides with the square of the tunneling wave function, and again there is no trace of the Hartle-Hawking wave function. This expression is valid in the whole interval from $\phi_e$ to $\phi_p$ and it correctly describes asymptotic behavior of ${P}_p(\phi,\chi)$ both at $\chi \sim \chi_e$ and at $\chi \sim \chi_p$. A similar investigation can be carried out for the theory $V(\phi) = V_o\ e^{\alpha\phi}$. The corresponding solution is \begin{eqnarray} {P}_p(\phi,\chi) &\sim & \exp\Bigl(-{3\over 8V(\chi)}\Bigr)\, \Bigl({1\over V(\chi)} - {1}\Bigr)\, \cdot \, \Bigl({1\over V(\phi)} - {1}\Bigr)\, V^{-1/2}(\phi)\ . \end{eqnarray} This expression gives a rather good approximation for ${P}_p(\phi,\chi)$ for all $\phi$ and $\chi$. The main result is that under certain conditions the properties of our Universe can be described by a time-independent probability distribution, which we have found for theories with polynomial and exponential effective potentials. Thus, inflation solves many problems of the big bang theory and ensures that this theory provides an excellent description of the local structure of the Universe. However, after making all kinds of improvements of this theory, we are now winding up with a model of a stationary Universe, in which the notion of the big bang loses its dominant position, being removed to the indefinite past. \section{\label{predictions} Predictions in quantum cosmology} \subsection{Moderate approach: comparing probabilities within the same Universe} When inflationary theory was first formulated, we did not know how much it was going to influence our understanding of the structure of the Universe. We were happy that inflation provided an easy explanation of the homogeneity of the Universe. However, we did not know that the same theory simultaneously predicts that on the extremely large scale the Universe becomes entirely inhomogeneous, and that this inhomogeneity is good, since it is one of the manifestations of the process of self-reproduction of inflationary Universe. The new picture of the Universe which emerges now is very unusual, and we are still in the process of learning how to ask proper questions in the context of the new cosmological paradigm. Previously we assumed that we live in a Universe which has the same properties everywhere (``cosmological principle''). Then one could make a guess about the most natural initial conditions in the Universe, and all the rest followed almost automatically. Now we learned that even if one begins with a non-uniform Universe, later it becomes extremely homogeneous on a very large scale. However, simultaneously it becomes absolutely non-uniform on a much greater scale. the Universe becomes divided into different exponentially large regions where the laws of low-energy physics can be different. In certain cases the relative fraction of volume of the Universe in a state with given fields or with a given density does not depend on time, whereas the total volume of all parts of the Universe continues growing exponentially. This change of the picture of the world is important by itself. However, it would be even better if we could use it to make certain predictions based on this picture. In this situation the problem of introduction of a proper measure of probability becomes most important. One of the most natural choices of such measure is given by the probability distribution $ P_p(\phi,\chi,t)$. The hypothesis behind this proposal is that we are typical, and therefore we live in those parts of the Universe where most other people do. The total number of people which can live in domains with given properties should be proportional to the total volume of these domains. There are two versions of this hypothesis, the moderate and the radical ones. The moderate version is based on investigation of $P_p(\phi,\chi,t)$ \cite{b19,LLM,GBLL}. If this distribution is stationary, then it seems reasonable to use it as a measure of the total volume of domains with any particular properties at any given moment of time $t$. The first example of this approach is given by the consideration of the axion problem. In the non-inflationary cosmology it was shown that the axion mass should be greater than $10^{-5}$ eV in order to avoid having too much energy stored in the axion field \cite{axion}. However, the derivation of this constraint fails in inflationary cosmology if one takes into account quantum fluctuations of the axion field and eternal production of domains where this field takes all its possible values. Then it can be shown that life of our type can appear only in those domains where the axion field is sufficiently small and under certain conditions discussed in \cite{LinAx} the standard constraint $m_a >10^{-5}$ eV disappears. Another interesting example is given by the probability distribution for finding the most probable values of the effective gravitational constant in the Brans-Dicke inflationary cosmology \cite{GBLL}. We have shown there that inflation in the Brans-Dicke theory leads to division of the Universe into different exponentially large domains with different values of the gravitational constant, and, correspondingly, with different values of density perturbations. Then one can use the probability distribution $P_p(\phi,\chi,t)$ to find most probable values of the gravitational constant. In this approach it is possible either to explain the anomalously large value of the Planck mass, or at least relate it to certain small parameters in the theory, e.g. to the small anisotropy of the microwave background radiation. Note, that the very language which we are using may sound somewhat strange. Indeed, typically the purpose is to express the anisotropy of the microwave background radiation via some fundamental parameters of the theory. In our case the Planck mass is not fundamental, and its value is anomalously large in those domains where the microwave background radiation is anomalously small. In what follows I will briefly describe some nonperturbative effects which may lead to a considerable local deviation of density from the critical density of a flat Universe \cite{OPEN}. Let us consider all parts of inflationary universe which contain a given field $\phi$ at a given moment of time $t$. One may wonder, what was the value of this field in those domains at the moment $t - H^{-1}$ ? The answer is simple: One should add to $\phi$ the value of its classical drift $\Delta \phi$ during the time $H^{-1}$, $\Delta \phi = \dot\phi H^{-1}$. One should also add the amplitude of a quantum jump $\delta \phi$. The typical jump is given by $\delta \phi = \pm {H\over 2\pi}$. At the last stages of inflation this quantity is by many orders of magnitude smaller than $\Delta \phi$. However, in which sense jumps $\pm {H\over 2\pi}$ are typical? If we consider any particular initial value of the field $\phi$, then the typical jump from this point is indeed given by $\pm {H\over 2\pi}$. However, if we are considering all domains with a given $\phi$ and trying to find all those domains from which the field $\phi$ could originate back in time, the answer may be quite different. Indeed, the total volume of all domains with a given field $\phi$ at any moment of time $t$ strongly depends on $\phi$: \,${P}_p(\phi) \sim \phi^{\sqrt{6\pi\over \lambda}\lambda_1} \sim \phi^{10^8}$, see eq. (\ref{SMALLPHI}). This means that the total volume of all domains which could jump towards the given field $\phi$ from the value $\phi +\delta \phi$ will be enhanced by a large additional factor $ {{P}_p(\phi +\delta \phi)\over {P}_p(\phi)} \sim \Bigl(1+{\delta\phi\over \phi}\Bigr)^{\sqrt{6\pi\over \lambda}\lambda_1}$. On the other hand, the probability of large jumps $\delta\phi$ is suppressed by the Gaussian factor $\exp\Bigl(-{2\pi^2\delta\phi^2\over H^2}\Bigr)$. One can easily verify that the product of these two factors has a sharp maximum at $\delta\phi = \lambda_1 \phi \cdot {H\over 2\pi}$, and the width of this maximum is of the order ${H\over 2\pi}$. In other words, most of the domains of a given field $\phi$ are formed due to jumps which are greater than the ``typical'' ones by a factor $\lambda_1 \phi \pm O(1) $. Our part of the Universe in the inflationary scenario with $V(\phi) = {\lambda\over 4} \phi^4$ is formed at $\phi~\sim~5$ (in the units $M_{\rm P} = 1$), and the constant $\lambda_1 \approx 2\sqrt{6\pi} \sim 8.68$ for our choice of boundary conditions. This means that our part of the Universe should be created as a result of a jump down which is about $\lambda_1 \phi \sim 40$ times greater than the standard jump. The standard jumps lead to density perturbations of the amplitude $\delta\rho \sim 5\cdot10^{-5} \rho_c$ (in the normalization of \cite{MyBook}). Thus, according to our nonperturbative analysis, we should live inside a region where density is smaller than the critical density by about $\delta\rho \sim 2\cdot10^{-3} \rho_c$. As we already mentioned, the probability of such fluctuations should be suppressed by $\exp\Bigl(-{2\pi^2\delta\phi^2\over H^2}\Bigr)$, which in our case gives the suppression factor $ \sim \exp(-10^3)$. It is well known that exponentially suppressed perturbations typically give rise to spherically symmetric bubbles. Note also, that the Gaussian distribution suppressing the amplitude of the perturbations refers to the amplitude of a perturbation in its maximum. It is possible that we live not in the place corresponding to the maximum of the fluctuation. However, this could only happen if the nonperturbative jump down was even greater in the amplitude that we expected. Meanwhile, as we already mentioned, the distribution of the amplitudes of such jumps has width of only about ${H\over 2\pi}$. This means that we should live very close to the center of the giant fluctuation, and the difference of energy densities between the place where we live and the center of the ``bubble'' should be only about the same amplitude as the typical perturbative fluctuation $\delta\rho \sim 5\cdot10^{-5} \rho_c$. In other words, we should live very close to the center of the nearly perfect spherically symmetric bubble, which contains matter with a smaller energy density than the matter outside it. It is very tempting to interpret this effect in such a way that the Universe around us becomes locally open. The true description of this effect is, of course, much more complicated; perhaps we should see the Hubble constant decreasing at large distances. This effect is extremely unusual. We became partially satisfied by our understanding of this effect only after we confirmed its existence by four different methods, including computer simulations \cite{OPEN}. However, it may happen that what we have found is simply a mathematical property of some particular hypersurfaces in inflationary universe, and it does not have any implications for the part of the Universe where we live. Indeed, it is quite legitimate to use the distributions like $P_p$ for descriptions of the structure of inflationary universe. However, it is not quite clear whether one can use them to evaluate probabilities. For example, instead of using the distribution $P_p(\phi,\chi,t)$ one may use the distribution $P_p(\phi,\chi \tau)$, where $\tau \sim \log a$, and many of our result (though not all of them) will change dramatically \cite{LLM,GBLL}. Still another answer will be obtained if one uses some other cut-off procedure, see \cite{VilNew}. The source of this ambiguity can be easily understood. The total volume of all parts of an eternally inflating Universe is infinite in the limit $t \to \infty$ (or $\tau \to \infty$). Therefore when we are trying to compare volumes of domains with different properties, we are comparing infinities. This leads to answers depending on the way we are making this comparison. It is possible that eventually we will resolve this problem. Still it will not guarantee that we are on the right track. Our use of $P_p$ as a probability measure was based on two hidden assumptions. The first assumption is that we are typical observers. The second assumption is that the number of typical observers is directly proportional to the volume of the Universe. If this is correct, then we should live in the place where most observers live, which should correspond to a maximum of $P_p$. However, is it absolutely clear that the probability for an observer to be born in a particular part of the Universe is directly proportional to its volume, or one should take into account something else? One cannot get any crop even from a very large field without having seeds first. The idea that life appears automatically once there is enough space to be populated may be too primitive. It is based on the assumption that one can describe emergence of life solely in terms of physics. It is certainly a most economical approach, and one should try to go as far as possible without invoking additional hypotheses. However, one should keep in mind that this approach may happen to be incomplete, especially if consciousness has its own degrees of freedom \cite{MyBook,Page}. Another related question is whether we are actually typical? Does it make any sense for each of us to calculate {\it a posteriori} what was the probability to be born Russian, Italian or Chinese? Should we insist on our own mediocrity, or, {\it vice versa}, should we try to explain why are we so special? After all, for a long time we thought that we had the aristocratic privilege to be the most intelligent species in the Universe. This, of course, may be wrong. Still, before using probabilities to calculate the likelihood of our existence in a particular part of the Universe, it may be a good idea to learn more about ourselves. I would take a certain risk to make a conjecture that until we understand what is our life and what is the nature of consciousness our understanding of quantum cosmology will remain fundamentally incomplete. \subsection{A more radical approach: comparing Universes with different coupling constants} Previously we compared volume of the parts of the Universe with some particular properties within one Universe. A more ambiguous program is based on a combination of the baby Universe theory and stochastic approach to inflation. The idea is that the coupling constants may take different values in different Universes, or, more precisely, in different quantum states of the Universe \cite{Coleman}. If this is the case, then perhaps we should live in those Universes where conditions are better and the total volume suitable for life is greater \cite{LinCosm}--\cite{GBL}. The total volume is given by ${\cal V}(\phi,\chi,t) = e^{ d(\lambda) H_{\rm \max}(\lambda)t}\, P_p(\phi,\chi,t)$. The first term in this expression is especially important. If (and this is a big ``if''!) one can compare the volumes of different Universes with different coupling constants at the same time $t$, the greatest volume will be occupied by the Universes with the largest product $d(\lambda) H_{\rm \max}(\lambda)$. For stationary $P_p(\phi,\chi,t) = P_p(\phi,\chi)$ the exponential growth of ${\cal V}(\phi,\chi,t)$ in the state with the largest $d(\lambda) H_{\rm \max}(\lambda)$ eventually beats all anthropic considerations. This may lead to a very sharp prediction of the coupling constants which maximize $d(\lambda) H_{\rm \max}(\lambda)$. Unfortunately, this immediately leads to a trouble. For example, in our investigation of the theory ${\lambda\over 4}\phi^4$ we have found that $H_{\rm max}= 2\sqrt{2\pi\over 3}$ does not depend on $\lambda$, whereas the fractal dimension $d(\lambda)$ has its maximum $d = 3$ in the limit $\lambda = 0$. This is a rather general conclusion which seems to suggest that the inflationary effective potential should be absolutely flat. But then there will be no density perturbations which are necessary for galaxy formation. One may try to avoid the problem with density perturbations assuming that they will be produced by cosmic strings \cite{Vil,Al}, but in the theory with absolutely flat potentials there will be no reheating and no cosmic strings. One may argue that this means that the potential should be {\it almost} flat, i.e. that it should be curved just enough to allow baryons and strings to be produced and life to appear. In fact, in such a case strings are not necessary. For example, one may consider the hybrid inflation model \cite{Hybrid}. In this model one can have good inflation and sufficiently large density perturbations without any need for cosmic strings even if the potential is extremely (though not exactly) flat. But the problem is that the gain in $e^{ d(\lambda) H_{\rm \max}(\lambda)t}$ eventually always beats the anthropic considerations, which pushes us towards the models with {\it exactly} flat potentials. If the effective potential is exactly flat, we have no reheating and no regular density perturbations, but even in this case life may appear in an infinite empty Universe with a very small but finite probability due to extremely improbable quantum fluctuations. Even though such conditions are extremely improbable, eventually we will be compensated by the indefinitely large growth of volume due to the term $e^{ d(\lambda) H_{\rm \max}(\lambda)t}$. However, in such a scenario there is no reason for our part of the Universe to be homogeneous on the scale $10^{28}$ cm, which is much greater than what is needed for our existence. One may also argue that if quantum cosmology pushes us outside of the limits of our normal existence, it probably puts us at the verge of being immediately extinct. Another example is related to the cosmological constant problem. Adding it to the Lagrangian also tends to increase $d(\lambda)$. Thus the considerations based on the investigation of the factor $e^{ d(\lambda) H_{\rm \max}(\lambda)t}$ may push us towards very large values of the vacuum energy density \cite{Vil}. Of course, one cannot go too far since our life cannot exist if the vacuum energy density is too large. However, anthropic considerations allow vacuum energy density $V_0$ two orders of magnitude greater than the critical density $\sim 10^{-29}$g$\cdot$cm$^{-3}$, i.e. two orders of magnitude greater than the present observational constraints on $V_0$ \cite{Weinberg82}. Moreover, as we just mentioned, the rapidly growing factor $e^{ d(\lambda) H_{\rm \max}(\lambda)t}$ should beat all anthropic considerations and should push $V_0$ even higher, which would be in a definite contradiction with the observational data. This indicates that something should be modified either in the radical approach described in this subsection or in our choice of the theories to which we applied this approach \cite{GBL}. Each of these possibilities can be true. First of all, it is not quite clear whether it makes any sense at all to compare volume of different Universes (rather than volume of different parts of the same Universe) at the same time. Then, in certain theories the probability distribution ${\cal V}(\phi,\chi,t)$ is not stationary \cite{GBLL}, so it cannot be represented as $e^{ d(\lambda) H_{\rm \max}(\lambda)t}\, P_p(\phi,\chi)$. Finally, under certain conditions the fastest growth of ${\cal V}(\phi,\chi,t)$ appears in the theories where the effective potential is not flat and the cosmological constant is not large \cite{GBL}. For example, adding a positive cosmological constant in the Starobinsky model {\it decreases} the rate of expansion. This pushes the cosmological constant to zero \cite{GBL}. Unfortunately, this cannot be considered as a possible solution of the cosmological constant problem since the same mechanism may push the cosmological constant even further, toward its negative values. To solve the cosmological constant problem it would be necessary to find a mechanism which pushes it to zero from both sides. It would be premature to make any final conclusions about the radical approach described above. The idea to use stochastic approach to inflation in order to understand our place in the world is extremely attractive. However, this powerful weapon should be used with caution, especially when one tries to extend its limits of applicability and use it in the context of the baby Universe theory. A possible attitude towards this approach is to consider it as a kind of ``theoretical experiment.'' We may try to use probabilistic considerations in our trial-and-error approach to quantum cosmology. If we get unreasonable results, this may serve as an indication that we are using quantum cosmology incorrectly. However, if we solve some problems which could not be solved in any other way, then we will have a reason to believe that we are moving in the right direction. In our opinion, at the present moment we do not have sufficient reasons to believe that the effective potential should be exactly flat, that the density perturbations should be produced by strings appearing after inflation, and that the cosmological constant should be as large as possible. On the other hand, it is not excluded that the stochastic approach to inflation, or some of its generalizations, will help us to solve the cosmological constant problem. This possibility certainly deserves further investigation. We will return to the possibility of making predictions and calculating probabilities in quantum cosmology in the next section, where we will consider the model of an open inflationary universe. \ \section{Inflation with $\Omega \not = 1$} \subsection{Inflation and flatness of the Universe} One of the most robust predictions of inflationary cosmology is that the Universe after inflation becomes extremely flat, which corresponds to $\Omega = 1$. Here $\Omega = {\rho\over \rho_c}$,\, $\rho_c$ being the energy density of a flat Universe. There were many good reasons to believe that this prediction was quite generic. The only way to avoid this conclusion is to assume that the Universe inflated only by about $e^{60}$ times. Exact value of the number of e-foldings $N$ depends on details of the theory and may somewhat differ from 60. It is important, however, that in any particular theory inflation by extra 2 or 3 e-foldings would make the Universe with $\Omega = 0.5$ or with $\Omega = 1.5$ almost exactly flat. Meanwhile, the typical number of e-foldings in chaotic inflation scenario in the theory ${m^2\over 2} \phi^2$ is not 60 but rather $10^{12}$. One can construct models where inflation leads to expansion of the Universe by the factor $e^{60}$. However, in most of such models small number of e-foldings simultaneously implies that density perturbations are extremely large. It may be possible to overcome this obstacle by a specific choice of the effective potential. However, this would be only a partial solution. If the Universe does not inflate long enough to become flat, then by the same token it does not inflate long enough to become homogeneous and isotropic. Thus, the main reason why it is difficult to construct inflationary models with $\Omega \not = 1$ is not the issue of fine tuning of the parameters of the models, which is necessary to obtain the Universe inflating exactly $e^{60}$ times, but the problem of obtaining a homogeneous Universe after inflation. Fortunately, it is possible to solve this problem, both for a closed Universe \cite{Lab} and for an open one \cite{Gott}--\cite{Arthur}. The main idea is to use the well known fact that the region of space created in the process of a quantum tunneling tends to have a spherically symmetric shape, and homogeneous interior, if the tunneling process is suppressed strongly enough. Then such bubbles of a new phase tend to evolve (expand) in a spherically symmetric fashion. Thus, if one could associate the whole visible part of the Universe with an interior of one such region, one would solve the homogeneity problem, and then all other problems will be solved by the subsequent relatively short stage of inflation. For a closed Universe the realization of this program is relatively straightforward \cite{Lab,Omega}. One should consider the process of quantum creation of a closed inflationary universe from ``nothing.'' If the probability of such a process is exponentially suppressed (and this is indeed the case if inflation is possible only at the energy density much smaller than the Planck density \cite{Creation}), then the Universe created that way will be rather homogeneous from the very beginning. The situation with an open Universe is much more complicated. Indeed, an open Universe is infinite, and it may seem impossible to create an infinite Universe by a tunneling process. Fortunately, this is not the case: any bubble formed in the process of the false vacuum decay looks from inside like an infinite open Universe \cite{CL}. If this Universe continues inflating inside the bubble \cite{Gott}--\cite{Arthur}, then we obtain an open inflationary Universe. These possibilities became a subject of an active investigation only very recently, and there are still many questions to be addressed. First of all, the bubbles created by tunneling are not {\it absolutely} uniform even if the probability of tunneling is very small. This may easily spoil the whole scenario since in the end of the day we need to explain why the microwave background radiation is isotropic with an accuracy of about $10^{-5}$. Previously we did not care much about initial homogeneities, but if the stage of inflation is short, we will the see original inhomogeneities imprinted in the perturbations of the microwave background radiation. The second problem is to construct realistic inflationary models where all these ideas could be realized in a natural way. Whereas for the closed Universe this problem can be easily solved \cite{Lab,Omega}, for an open Universe we again meet complications. It would be very nice to to obtain an open Universe in a theory of just one scalar field \cite{BGT}. However, in practice it is not very easy to obtain a satisfactory model of this type. Typically one is forced either to introduce very complicated effective potentials, or consider theories with nonminimal kinetic terms for the inflaton field \cite{Bucher}. This makes the models not only fine-tuned, but also rather complicated. It is very good to know that the models of such type in principle can be constructed, but it is also very tempting to find a more natural realization of the inflationary universe scenario which would give inflation with $\Omega < 1$. This goal can be achieved if one considers models of two scalar fields \cite{Omega}. One of them may be the standard inflaton field $\phi$ with a relatively small mass, another may be, e.g., the scalar field responsible for the symmetry breaking in GUTs. The presence of two scalar fields allows one to obtain the required bending of the inflaton potential by simply changing the definition of the inflaton field in the process of inflation. At the first stage the role of the inflaton is played by a heavy field with a steep barrier in its potential, while on the second stage the role of the inflaton is played by a light field, rolling in a flat direction ``orthogonal'' to the direction of quantum tunneling. This change of the direction of evolution in the space of scalar fields removes the naturalness constraints for the form of the potential, which are present in the case of one field. Inflationary models of this type are quite simple, yet they have many interesting features. In these models the Universe consists of infinitely many expanding bubbles immersed into exponentially expanding false vacuum state. Each of these bubbles inside looks like an open Universe, but the values of $\Omega$ in these Universes may take any value from $1$ to $0$. In some of these models the situation is even more complicated: Interior of each bubble looks like an infinite Universe with an effective value of $\Omega$ slowly decreasing to $\Omega = 0$ at an exponentially large distance from the center of the bubble. We will call such Universes quasiopen. Thus, rather unexpectedly, we are obtaining a large variety of interesting and previously unexplored possibilities. Our discussion of these possibilities will follow our recent paper with Arthur Mezhlumian \cite{Arthur}. \subsection{\label{Bubbles} Tunneling probability and spherical symmetry} Typically it is assumed that the bubbles containing open Universes are exactly spherically symmetric (or, to be more accurate, $O(3,1)$-symmetric \cite{CL}). Meanwhile in realistic situations this condition may be violated for several reasons. First of all, the bubble may be formed not quite symmetric. Then its shape may change even further due to growth of its initial inhomogeneities and due to quantum fluctuations which appear during the bubble wall expansion. As we will see, this may cause a lot of problems if one wishes to maintain the degree of anisotropy of the microwave background radiation inside the bubble at the level of $10^{-5}$. First of all, let us consider the issue of symmetry of a bubble at the moment of its formation. For simplicity we will investigate the models where tunneling can be described in the thin wall approximation. We will neglect gravitational effects, which is possible as far as the initial radius $r$ of the bubble is much smaller than $H^{-1}$. In this approximation (which works rather well for the models to be discussed) euclidean action of the $O(4)$-symmetric instanton describing bubble formation is given by \begin{equation}\label{o6} S = - {\epsilon\over 2} \pi^2 r^4 + 2\pi^2 r^3 s \ . \end{equation} Here $r$ is the radius of the bubble at the moment of its formation, $\epsilon$ is the difference of $V(\phi)$ between the false vacuum $\phi_{\rm initial}$ and the true vacuum $\phi_{\rm final}$, and $s$ is the surface tension, \begin{equation}\label{o7} s = \, \int_{\phi_{\rm initial}}^{\phi_{\rm final}} \sqrt{ 2(V(\phi) - V(\phi_{\rm final}))}\, d\phi \ . \end{equation} The radius of the bubble can be obtained from the extremum of (\ref{o6}) with respect to $r$: \begin{equation}\label{o8} r = {3s\over \epsilon } \ . \end{equation} Let us check how the action $S $ will change if one consider a bubble of a radius $r + \Delta r$. Since the first derivative of $S $ at its extremum vanishes, the change will be determined by its second derivative, \begin{equation}\label{09} \Delta S = {1\over 2} S'' (\Delta r)^2 = 9\pi^2\, {s^2\over \epsilon}\, (\Delta r)^2 \ . \end{equation} Now we should remember that all trajectories which have an action different from the action at extremum by no more than $1$ are quite legitimate. Thus the typical deviation of the radius of the bubble from its classical value (\ref{o8}) can be estimated from the condition $\Delta S \sim 1$, which gives \begin{equation}\label{o10} |\Delta r| \sim {\sqrt\epsilon\over 3\pi \,s} \ . \end{equation} Note, that even though we considered spherically symmetric perturbations, our estimate is based on corrections proportional to $(\delta r)^2$, and therefore it should remain valid for perturbations which have an amplitude $\Delta r$, but change their sign in different parts of the bubble surface. Thus, eq. (\ref{o10}) gives an estimate of a typical degree of asymmetry of the bubble at the moment of its creation: \begin{equation}\label{o11} A(r) \equiv {|\Delta r| \over r} \sim {\epsilon\sqrt\epsilon\over 3\pi \,s^2} \ . \end{equation} This simple estimate exactly coincides with the corresponding result obtained by Garriga and Vilenkin \cite{VilGarr} in their study of quantum fluctuations of bubble walls. It was shown in \cite{VilGarr} that when an empty bubble begins expanding, the typical deviation $\Delta r$ remains constant. Therefore the asymmetry given by the ratio ${|\Delta r| \over r}$ gradually vanishes. This is a pretty general result: Waves produced by a brick falling to a pond do not have the shape of a brick, but gradually become circles. However, in our case the situation is somewhat more complicated. The wavefront produced by a brick in inflationary background preserves the shape of the brick if its size is much greater than $H^{-1}$. Indeed, the wavefront moves with the speed approaching the speed of light, whereas the distance between different parts of a region with initial size greater than $H^{-1}$ grows with a much greater (and ever increasing) speed. This means that inflation stretches the wavefront without changing its shape on scale much greater than $H^{-1}$. Therefore during inflation which continues inside the bubble the symmetrization of its shape occurs only in the very beginning, until the radius of the bubble approaches $H^{-1}$. At this first stage expansion of the bubble occurs mainly due to the motion of the walls rather than due to inflationary stretching of the Universe, and our estimate of the bubble wall asymmetry as well as the results obtained by Garriga and Vilenkin for the empty bubble remain valid. At the moment when the radius of the bubble becomes equal to $H^{-1}$ its asymmetry becomes \begin{equation}\label{o12} A(H^{-1}) \sim {|\Delta r| H} \sim {\sqrt\epsilon H\over 3\pi \,s} \ , \end{equation} and the subsequent expansion of the bubble does not change this value very much. Note that the Hubble constant here is determined by the vacuum energy {\it after} the tunneling, which may differ from the initial energy density $\epsilon$. The deviation of the shape of the bubble from spherical symmetry implies that the beginning of the second stage of inflation inside the bubble will be not exactly synchronous, with the delay time $\Delta t \sim \Delta r$. This, as usual, may lead to adiabatic density perturbations on the horizon scale of the order of $H\Delta t$, which coincides with the bubble asymmetry $A$ after its size becomes greater than $H^{-1}$, see Eq.\ (\ref{o12}). To estimate this contribution to density perturbations, let us consider again the simplest model with the effective potential \begin{equation}\label{o1} V(\phi) = {m^2\over 2} \phi^2 - {\delta\over 3} \phi^3 + {\lambda\over 4}\phi^4 \ . \end{equation} Now we will consider it in the limit where the two minima of this potential have almost the same depth, which is necessary for validity of the thin wall approximation. In this case $2\delta^2 = 9 M^2\lambda$, and the effective potential (\ref{o1}) looks approximately like ${\lambda \over 4} \phi^2 (\phi - \phi_0)^2$, where $\phi_0 = {2\delta\over 3\lambda} = \sqrt {2\over \lambda} {M}$ is the position of the local minimum of the effective potential. The surface tension in this model is given by $s = \sqrt{\lambda\over 2} {\phi_0^3\over 6} = { M^3\over 3\lambda}$ \cite{Tunn}. We will also introduce a phenomenological parameter $\mu$, such that $\mu {M^4\over 16\lambda} = \epsilon$. The smallness of this parameter controls applicability of the thin-wall approximation, since the value of the effective potential near the top of the potential barrier at $\phi = \phi_0/2$ is given by $M^4\over 16\lambda$. Then our estimate of density perturbations associated with the bubble wall (\ref{o12}) gives \begin{equation}\label{o14} \left. {\delta\rho\over \rho}\right|_{\rm bubble} \sim A(H^{-1}) \sim {\sqrt{\mu\lambda} H\over 4 \pi M} \ . \end{equation} Here $H$ is the value of the Hubble constant at the beginning of inflation inside the bubble. In order to have $\left. {\delta\rho\over \rho}\right|_{\rm bubble} {\ \lower-1.2pt\vbox{\hbox{\rlap{$<$}\lower5pt\vbox{\hbox{$\sim$}}}}\ } 5 \times 10^{-5}$ (the number $5 \times 10^{-5}$ corresponds to the amplitude of density perturbations in the COBE normalization) one should have \begin{equation}\label{o17} \left. {\delta\rho\over \rho}\right|_{\rm bubble} \sim \, {\sqrt{\mu\lambda} H\over 4 \pi M}\, {\ \lower-1.2pt\vbox{\hbox{\rlap{$<$}\lower5pt\vbox{\hbox{$\sim$}}}}\ }\, 5 \times 10^{-5} \ . \end{equation} For $H\ll M$ perturbations produced by the bubble walls may be sufficiently small even if the coupling constants are relatively large and the bubbles at the moment of their formation are very inhomogeneous. There is a long way from our simple estimates to the full theory of anisotropies of cosmic microwave background induced by fluctuations of the domain wall. In particular, the significance of this effect will clearly depend on the value of $\Omega$ \cite{Open}. The constraint (\ref{o17}) may appear only if one can ``see'' the scale at which the bubble walls have imprinted their fluctuations. If inflation is long enough, this scale becomes exponentially large, we do not see the fluctuations due to bubble walls, but then we return to the standard inflationary scenario of a flat inflationary universe. However, for $\Omega \ll 1$ inflation is short, and it does not preclude us from seeing perturbations in a vicinity of the bubble walls. In such a case one should take the constraint (\ref{o17}) very seriously. \subsection{\label{Simplest} The simplest model of a (quasi)open inflationary Universe} In this section we will explore an extremely simple model of two scalar fields, where the Universe after inflation becomes open (or quasiopen, see below) in a very natural way \cite{Omega}. Consider a model of two noninteracting scalar fields, $\phi$ and $\sigma$, with the effective potential \begin{equation}\label{3} V(\phi, \sigma) = {m^2\over 2}\phi^2 + V(\sigma) \ . \end{equation} Here $\phi$ is a weakly interacting inflaton field, and $\sigma$, for example, can be the field responsible for the symmetry breaking in GUTs. We will assume that $V(\sigma)$ has a local minimum at $\sigma = 0$, and a global minimum at $\sigma_0 \not = 0$, just as in the old inflationary theory. For definiteness, we will assume that this potential is given by ${M^2\over 2} \sigma^2 - {\alpha M } \sigma^3 + {\lambda\over 4}\sigma^4 + V(0)$, with $V(0) \sim {M^4\over 4 \lambda}$, but it is not essential; no fine tuning of the shape of this potential will be required. Note that so far we did not make any unreasonable complications to the standard chaotic inflation scenario; at large $\phi$ inflation is driven by the field $\phi$, and the GUT potential is necessary in the theory anyway. In order to obtain density perturbations of the necessary amplitude the mass $m$ of the scalar field $\phi$ should be of the order of $10^{-6} M_{\rm P} \sim 10^{13}$ GeV \cite{MyBook}. Inflation begins at $V(\phi, \sigma) \sim M_{\rm P}^4$. At this stage fluctuations of both fields are very strong, and the Universe enters the stage of self-reproduction, which finishes for the field $\phi$ only when it becomes smaller than $M_{\rm P} \sqrt{M_{\rm P}\over m}$ and the energy density drops down to $m M_{\rm P}^3 \sim 10^{-6} M_{\rm P}^4$ \cite{MyBook}. Quantum fluctuations of the field $\sigma$ in some parts of the Universe put it directly to the absolute minimum of $V(\sigma)$, but in some other parts the scalar field $\sigma$ appears in the local minimum of $V(\sigma)$ at $\sigma = 0$. We will follow evolution of such domains. Since the energy density in such domains will be greater, their volume will grow with a greater speed, and therefore they will be especially important for us. One may worry that all domains with $\sigma = 0$ will tunnel to the minimum of $V(\sigma)$ at the stage when the field $\phi$ was very large and quantum fluctuations of the both fields were large too. This may happen if the Hubble constant induced by the scalar field $\phi$ is much greater than the curvature of the potential $V(\sigma)$: \begin{equation}\label{s1} {m\phi\over M_{\rm P}} {\ \lower-1.2pt\vbox{\hbox{\rlap{$>$}\lower5pt\vbox{\hbox{$\sim$}}}}\ } M \ . \end{equation} This decay can be easily suppressed if one introduces a small interaction $g^2\phi^2\sigma^2$ between these two fields, which stabilizes the state with $\sigma = 0$ at large $\phi$. Another possibility is to add a nonminimal interaction with gravity of the form $-{\xi\over 2} R\phi^2$, which makes inflation impossible for $\phi > {M_{\rm P}\over 8\phi\xi}$. In this case the condition (\ref{s1}) will never be satisfied. However, there is a much simpler answer to this worry. If the effective potential of the field $\phi$ is so large that the field $\sigma$ can easily jump to the true minimum of $V(\sigma)$, then the Universe becomes divided into infinitely many domains with all possible values of $\sigma$ distributed in the following way \cite{Star,MyBook}: \begin{equation}\label{s2} {P(\sigma= 0)\over P(\sigma = \sigma_0)} \sim \exp\left({3M^4_{\rm P}\over 8 V(\phi,0)} - {3M^4_{\rm P}\over 8V(\phi,\sigma)}\right) = \exp\left({3M^4_{\rm P}\over 4(m^2\phi^2 + 2V(0))} - {3M^4_{\rm P}\over 4 m^2\phi^2}\right)\ . \end{equation} One can easily check that at the moment when the field $\phi$ decreases to ${M M_{\rm P}\over m}$ and the condition (\ref{s1}) becomes violated, we will have \begin{equation}\label{s3} {P(0)\over P(\sigma_0)} \sim \exp\left(-{C\over \lambda}\right) \ , \end{equation} where $C$ is some constant, $C = O(1)$. After this moment the probability of the false vacuum decay typically becomes much smaller. Thus the fraction of space which survives in the false vacuum state $\sigma = 0$ until this time typically is very small, but finite (and calculable). It is important, that these rare domains with $\sigma = 0$ eventually will dominate the volume of the Universe since if the probability of the false vacuum decay is small enough, the volume of the domains in the false vacuum will continue growing exponentially without end. The main idea of our scenario can be explained as follows. Because the fields $\sigma$ and $\phi$ do not interact with each other, and the dependence of the probability of tunneling on the vacuum energy at the GUT scale is negligibly small \cite{CL}, tunneling to the minimum of $V(\sigma)$ may occur with approximately equal probability at all sufficiently small values of the field $\phi$ (see, however, below). The parameters of the bubbles of the field $\sigma$ are determined by the mass scale $M$ corresponding to the effective potential $V(\sigma)$. This mass scale in our model is much greater than $m$. Thus the duration of tunneling in the Euclidean ``time'' is much smaller than $m^{-1}$. Therefore the field $\phi$ practically does not change its value during the tunneling. If the probability of decay at a given $\phi$ is small enough, then it does not destroy the whole vacuum state $\sigma = 0$ \cite{GW}; the bubbles of the new phase are produced all the way when the field $\phi$ rolls down to $\phi = 0$. In this process the Universe becomes filled with (nonoverlapping) bubbles immersed in the false vacuum state with $\sigma = 0$. Interior of each of these bubbles represents an open Universe. However, these bubbles contain {\it different} values of the field $\phi$, depending on the value of this field at the moment when the bubble formation occurred. If the field $\phi$ inside a bubble is smaller than $3 M_{\rm P}$, then the Universe inside this bubble will have a vanishingly small $\Omega$, at the age $10^{10}$ years after the end of inflation it will be practically empty, and life of our type would not exist there. If the field $\phi$ is much greater than $3 M_{\rm P}$, the Universe inside the bubble will be almost exactly flat, $\Omega = 1$, as in the simplest version of the chaotic inflation scenario. It is important, however, that {\it in an eternally existing self-reproducing Universe there will be infinitely many Universes containing any particular value of $\Omega$, from $\Omega = 0$ to $\Omega = 1$}, and one does not need any fine tuning of the effective potential to obtain a Universe with, say, $0.2 <\Omega < 0.3$ Of course, one can argue that we did not solve the problem of fine tuning, we just transformed it into the fact that only a very small percentage of all Universes will have $0.2 <\Omega < 0.3$. However, first of all, we achieved our goal in a very simple theory, which does not require any artificial potential bending and nonminimal kinetic terms. Then, there may be some reasons why it is preferable for us to live in a Universe with a small (but not vanishingly small) $\Omega$. The simplest way to approach this problem is to find how the probability for the bubble production depends on $\phi$. As we already pointed out, for small $\phi$ this dependence is not very strong. On the other hand, at large $\phi$ the probability rapidly grows and becomes quite large at $\phi > {M M_{\rm P}\over m}$. This may suggest that the bubble production typically occurs at $\phi > {M M_{\rm P}\over m}$, and then for ${M\over m} \gg 3$ we typically obtain flat Universes, $\Omega = 1$. This is another manifestation of the problem of premature decay of the state $\sigma = 0$ which we discussed above. Moreover, even if the probability to produce the Universes with different $\phi$ were entirely $\phi$-independent, one could argue that the main volume of the habitable parts of the Universe is contained in the bubbles with $\Omega = 1$, since the interior of each such bubble inflated longer. Indeed, the total volume of each bubble created in a state with the field $\phi$ during inflation in our model grows by the factor of $\exp{6\pi\phi^2\over M_{\rm P}^2}$ \cite{MyBook}. It seems clear that the bubbles with greater $\phi$ will give the largest contribution to the total volume of the Universe after inflation. This would be the simplest argument in favor of the standard prediction $\Omega = 1$ even in our class of models. However, there exist several ways of resolving this problem: involving coupling $g^2\phi^2\sigma^2$, which stabilizes the state $\sigma = 0$ at large $\phi$, or adding nonminimal interaction with gravity of the form $-{\xi\over 2} R\phi^2$. In either way one can easily suppress production of the Universes with $\Omega = 1$. Then the maximum of probability will correspond to some value $\Omega < 1$, which can be made equal to any given number from $1$ to $0$ by changing the parameters $g^2$ and $\xi$. For example, let us add to the Lagrangian the term $-{\xi\over 2} R\phi^2$. This term makes inflation impossible for $\phi > \phi_c = {M_{\rm P}\over \sqrt{8\pi\xi}}$. If initial value of the field $\phi$ is much smaller than $\phi_c$, the size of the Universe during inflation grows $\exp{2\pi\phi^2\over M_{\rm P}^2}$ times, and the volume grows $\exp{6\pi\phi^2\over M_{\rm P}^2}$ times, as in the theory ${m^2\over 2} \phi^2$ with $\xi = 0$. For initial $\phi$ approaching $\phi_c$ these expressions somewhat change, but in order to get a very rough estimate of the increase of the size of the Universe in this model (which is sufficient to get an illustration of our main idea) one can still use the old expression $\exp{2\pi\phi^2\over M_{\rm P}^2}$. This expression reaches its maximum near $\phi = \phi_c$, at which point the effective gravitational constant becomes infinitely large and inflationary regime ceases to exist \cite{Maeda,GBL}. Thus, one may argue that in this case the main part of the volume of the Universe will appear from the bubbles with initial value of the field $\phi$ close to $\phi_c$. For $\xi \ll 4.4\times 10^{-3}$ one has $\phi_c \gg 3 M_{\rm P}$. In this case one would have typical Universes expanding much more than $e^{60}$ times, and therefore $\Omega \approx 1$. For $\xi \gg 4.4\times 10^{-3}$ one has $\phi_c \ll 3 M_{\rm P}$, and therefore one would have $\Omega \ll 1$ in all inflationary bubbles. It is clear that by choosing particular values of the constant $\xi$ in the range of $\xi \sim 4.4\times 10^{-3}$ one can obtain the distribution of the Universes with the maximum of the distribution concentrated near any desirable value of $\Omega < 1$. Note that the position of the peak of the distribution is very sensitive to the value of $\xi$: to have the peak concentrated in the region $0.2 < \Omega < 0.3$ one would have to fix $\xi$ (i.e. $\phi_c$) with an accuracy of few percent. Thus, in this approach to the calculation of probabilities to live in a Universe with a given value of $\Omega$ we still have the problem of fine tuning. However, calculation of probabilities in the context of the theory of a self-reproducing Universe is a very ambiguous process, and it is even not quite clear that this process makes any sense at all. For example, we may formulate the problem in a different way. Consider a domain of the false vacuum with $\sigma = 0$ and $\phi = \phi_1$. After some evolution it produces one or many bubbles with $\sigma = \sigma_0$ and the field $\phi$ which after some time becomes equal to $\phi_2$. One may argue that the most efficient way this process may go is the way which in the end produces the greater volume. Indeed, for the inhabitants of a bubble it does not matter how much time did it take for this process to occur. The total number of observers produced by this process will depend on the total volume of the Universe at the hypersurface of a given density, i.e. on the hypersurface of a given $\phi$. If the domain instantaneously tunnels to the state $\sigma_0$ and $\phi_1$, and then the field $\phi$ in this domain slowly rolls from $\phi_1$ to $\phi_2$, then the volume of this domain grows $\exp \Bigl({2\pi\over M_{\rm P}^2} (\phi_1^2 -\phi_2^2)\Bigr)$ times \cite{MyBook}. Meanwhile, if the tunneling takes a long time, then the field $\phi$ rolls down extremely slowly being in the false vacuum state with $\sigma = 0$. In this state the Universe expands much faster than in the state with $\sigma = \sigma_0$. Since it expands much faster, and it takes the field much longer to roll from $\phi_1$ to $\phi_2$, the trajectories of this kind bring us much greater volume. This may serve as an argument that most of the volume is produced by the bubbles created at a very small $\phi$, which leads to the Universes with very small $\Omega$. One may use another set of considerations, studying all trajectories beginning at $\phi_1, t_1$ and ending at $\phi_2, t_2$. This will bring us another answer, or, to be more precise, another set of answers, which will depend on the choice of the time parametrization \cite{LLM}. Still another answer will be obtained by the method recently proposed by Vilenkin, who suggested to introduce a particular cutoff procedure which (almost) completely eliminates dependence of the final answer on the time parametrization \cite{VilNew}. A more radical possibility would be to integrate over all time parametrizations. This task is very complicated, but it would completely eliminate dependence of the final answer on the time parametrization. There is a very deep reason why the calculation of the probability to obtain a Universe with a given $\Omega$ is so ambiguous. We have discussed this reason in Sect. 3.1 in general terms; let us see how the situation looks in application to the open Universe scenario. For those who lives inside a bubble there is be no way to say at which stage (at which time from the point of view of an external observer) this bubble was produced. Therefore one should compare {\it all} of these bubbles produced at all possible times. The self-reproducing Universe should exist for indefinitely long time, and therefore it should contain infinitely many bubbles with all possible values of $\Omega$. Comparing infinities is a very ambiguous task, which gives results depending on the procedure of comparison. For example, one can consider an infinitely large box of apples and an infinitely large box of oranges. One may pick up one apple and one orange, then one apple and one orange, over and over again, and conclude that there is an equal number of apples and oranges. However, one may also pick up one apple and two oranges, and then one apple and two oranges again, and conclude that there is twice as many oranges as apples. The same situation happens when one tries to compare the number of bubbles with different values of $\Omega$. If we would know how to solve the problem of measure in quantum cosmology, perhaps we would be able to obtain something similar to an open Universe in the trivial $\lambda\phi^4$ theory without any first order phase transitions \cite{OPEN}, see Sect. 3.1. In the meantime, it is already encouraging that in our scenario there are infinitely many inflationary universes with all possible value of $\Omega < 1$. We can hardly live in the empty bubbles with $\Omega = 0$. As for the choice between the bubbles with different nonvanishing values of $\Omega < 1$, it is quite possible that eventually we will find out an unambiguous way of predicting the most probable value of $\Omega$, and we are going to continue our work in this direction. However, as we already discussed in the previous section, it might also happen that this question is as meaningless as the question whether it is more probable to be born as a Chinese rather than as an Italian. It is quite conceivable that the only way to find out in which of the bubbles do we live is to make observations. Some words of caution are in order here. The bubbles produced in our simple model are not {\it exactly} open Universes. Indeed, in the models discussed in \cite{CL}--\cite{BGT} the time of reheating (and the temperature of the Universe after the reheating) was synchronized with the value of the scalar field inside the bubble. In our case the situation is very similar, but not exactly. Suppose that the Hubble constant induced by $V(0)$ is much greater than the Hubble constant related to the energy density of the scalar field $\phi$. Then the speed of rolling of the scalar field $\phi$ sharply increases inside the bubble. Thus, in our case the field $\sigma$ synchronizes the motion of the field $\phi$, and then the hypersurface of a constant field $\phi$ determines the hypersurface of a constant temperature. In the models where the rolling of the field $\phi$ can occur only inside the bubble (we will discuss such a model shortly) the synchronization is precise, and everything goes as in the models of refs. \cite{CL}--\cite{BGT}. However, in our simple model the scalar field $\phi$ moves down outside the bubble as well, even though it does it very slowly. Thus, synchronization of motion of the fields $\sigma$ and $\phi$ is not precise; hypersurface of a constant $\sigma$ ceases to be a hypersurface of a constant density. For example, suppose that the field $\phi$ has taken some value $\phi_0$ near the bubble wall when the bubble was just formed. Then the bubble expands, and during this time the field $\phi$ outside the wall decreases, as $\exp \Bigl(-{m^2t\over 3 H_1}\Bigr)$, where $H_1 \approx H(\phi = \sigma = 0)$ is the Hubble constant at the first stage of inflation, $H_1 \approx \sqrt{8\pi V(0)\over 3 M_{\rm P}^2}$. At the moment when the bubble expands $e^{60}$ times, the field $\phi$ in the region just reached by the bubble wall decreases to $\phi_o\exp \Bigl(-{20 m^2\over H^2_1}\Bigr)$ from its original value $\phi_0$. the Universe inside the bubble is a homogeneous open Universe only if this change is negligibly small. This may not be a real problem. Indeed, let us assume that $V(0) ={\tilde M}^4$, where ${\tilde M} = 10^{17}$ GeV. (Typically the energy density scale $\tilde M$ is related to the particle mass as follows: ${\tilde M} \sim \lambda^{-1/4} M$.) In this case $H_1 = 1.7 \times 10^{15}$ GeV, and for $m = 10^{13}$ GeV one obtains ${20 m^2\over H_1^2} \sim 10^{-4}$. In such a case a typical degree of distortion of the picture of a homogeneous open Universe is very small. Still this issue requires careful investigation. When the bubble wall continues expanding even further, the scalar field outside of it eventually drops down to zero. Then there will be no new matter created near the wall. Instead of infinitely large homogeneous open Universes we are obtaining spherically symmetric islands of a size much greater than the size of the observable part of our Universe. We do not know whether this unusual picture is an advantage or a disadvantage of our model. Is it possible to consider different parts of the same exponentially large island as domains of different ``effective'' $\Omega$? Can we attribute some part of the dipole anisotropy of the microwave background radiation to the possibility that we live somewhere outside of the center of such island? In any case, as we already mentioned, in the limit $m^2 \ll H_1^2$ we do not expect that the small deviations of the geometry of space inside the bubble from the geometry of an open Universe can do much harm to our model. Our model admits many generalizations, and details of the scenario which we just discussed depend on the values of parameters. Let us forget for a moment about all complicated processes which occur when the field $\phi$ is rolling down to $\phi = 0$, since this part of the picture depends on the validity of our ideas about initial conditions. For example, there may be no self-reproduction of inflationary domains with large $\phi$ if one considers an effective potential of the field $\phi$ which is very curved at large $\phi$. However, there will be self-reproduction of the Universe in a state $\phi = \sigma = 0$, as in the old inflation scenario. Then the main portion of the volume of the Universe will be determined by the processes which occur when the fields $\phi$ and $\sigma$ stay at the local minimum of the effective potential, $\phi = \sigma = 0$. For definiteness we will assume here that $V(0) = {\tilde M}^4$, where ${\tilde M}$ is the stringy scale, ${\tilde M} \sim 10^{17} - 10^{18}$ GeV. Then the Hubble constant $H_1 = \sqrt{8\pi V(0)\over 3M^2_{\rm P}} \sim \sqrt{8\pi \over 3} {{\tilde M}^2\over M_{\rm P}}$ created by the energy density $V(0)$ is much greater than $m \sim 10^{13}$ GeV. In such a case the scalar field $\phi$ will not stay exactly at $\phi = 0$. It will be relatively homogeneous on the horizon scale $H_1^{-1}$, but otherwise it will be chaotically distributed with the dispersion $\langle\phi^2\rangle = {3H^4\over 8\pi^2m^2}$ \cite{MyBook}. This means that the field $\phi$ inside each of the bubbles produced by the decay of the false vacuum can take any value $\phi$ with the probability \begin{equation}\label{4} P \sim \exp\left(-{\phi^2\over 2 \langle\phi^2\rangle}\right) \sim \exp\left(-{3m^2 \phi^2M_{\rm P}^4\over 16 {\tilde M}^8}\right) \ . \end{equation} One can check that for ${\tilde M} \sim 4.3\times10^{17}$ GeV the typical value of the field $\phi$ inside the bubbles will be $\sim 3\times 10^{19}$ GeV. Thus, for ${\tilde M} > 4.3\times10^{17}$ GeV most of the Universes produced during the vacuum decay will be flat, for ${\tilde M} < 4.3\times10^{17}$ GeV most of them will be open. It is interesting that in this version of our model the percentage of open Universes is determined by the stringy scale (or by the GUT scale). However, since the process of bubble production in this scenario goes without end, the total number of Universes with any particular value of $\Omega < 1$ will be infinitely large for any value of ${\tilde M}$. Thus this model shows us is the simplest way to resurrect some of the ideas of the old inflationary theory with the help of chaotic inflation, and simultaneously to obtain inflationary Universe with $\Omega < 1$. Note that this version of our model will not suffer for the problem of incomplete synchronization. Indeed, the average value of the field $\phi$ in the false vacuum outside the bubble will remain constant until the bubble triggers its decrease. However, this model, just as its previous version, may suffer from another problem. The Hubble constant $H_1$ before the tunneling in this model was much greater than the Hubble constant $H_2$ at the beginning of the second stage of inflation. Therefore the fluctuations of the scalar field before the tunneling were very large, $\delta \phi \sim {H_1\over 2 \pi}$, much greater than the fluctuations generated after the tunneling, $\delta \phi \sim {H_2\over 2 \pi}$. This may lead to very large density perturbations on the scale comparable to the size of the bubble. For the models with $\Omega = 1$ this effect would not cause any problems since such perturbations would be far away over the present particle horizon, but for small $\Omega$ this may lead to unacceptable anisotropy of the microwave background radiation. Fortunately, this may not be a real difficulty. A possible solution is very similar to the bubble symmetrization described in the previous section. Indeed, let us consider more carefully how the long wave perturbations produced outside the bubble may penetrate into it. At the moment when the bubble is formed, it has a size (\ref{o8}), which is smaller than $H_1^{-1}$ \cite{CL}. Then the bubble walls begin moving with the speed gradually approaching the speed of light. At this stage the comoving size of the bubble (from the point of view of the original coordinate system in the false vacuum) grows like \begin{equation}\label{n1} r(t) = \int_{0}^{t}{dt e^{-H_1 t}} = H_1^{-1} (1 - e^{-H_1 t}) \ . \end{equation} During this time the fluctuations of the scalar field $\phi$ of the amplitude ${H_1\over 2\pi}$ and of the wavelength $H_1^{-1}$, which previously were outside the bubble, gradually become covered by it. When these perturbations are outside the bubble, inflation with the Hubble constant $H_1$ prevents them from oscillating and moving. However, once these perturbations penetrate inside the bubble, their amplitude becomes decreasing \cite{MZ,SP}. Indeed, since the wavelength of the perturbations is $\sim H_1^{-1} \ll H_2^{-1} \ll m^{-1}$, these perturbations move inside the bubbles as relativistic particles, their wavelength grow as $a(t)$, and their amplitude decreases just like an amplitude of electromagnetic field, $\delta\phi \sim a^{-1}(t)$, where $a$ is the scale factor of the Universe inside a bubble \cite{MZ}. This process continues until the wavelength of each perturbation reaches $H_2^{-1}$ (already at the second stage of inflation). During this time the wavelength grows ${H_1\over H_2}$ times, and the amplitude decreases ${H_2\over H_1}$ times, to become the standard amplitude of perturbations produced at the second stage of inflation: $ {H_2\over H_1}\, {H_1\over 2\pi} = {H_2\over 2\pi}$. In fact, one may argue that this computation was too naive, and that these perturbations should be neglected altogether. Typically we treat long wave perturbations in inflationary universe like classical wave for the reason that the waves with the wavelength much greater than the horizon can be interpreted as states with extremely large occupation numbers \cite{MyBook}. However, when the new born perturbations (i.e. fluctuations which did not acquire an exponentially large wavelength yet) enter the bubble (i.e. under the horizon), they effectively return to the realm of quantum fluctuations again. Then one may argue that one should simply forget about the waves with the wavelengths small enough to fit into the bubble, and consider perturbations created at the second stage of inflation not as a result of stretching of these waves, but as a new process of creation of perturbations of an amplitude ${H_2\over 2\pi}$. One may worry that perturbations which had wavelengths somewhat greater than $H_1^{-1}$ at the moment of the bubble formation cannot completely penetrate into the bubble. If, for example, the field $\phi$ differs from some constant by $+{H_1\over 2\pi}$ at the distance $H_1^{-1}$ to the left of the bubble at the moment of its formation, and by $-{H_1\over 2\pi}$ at the distance $H_1^{-1}$ to the right of the bubble, then this difference remains frozen independently of all processes inside the bubble. This may suggest that there is some unavoidable asymmetry of the distribution of the field inside the bubble. However, the field inside the bubble will not be distributed like a straight line slowly rising from $-{H_1\over 2\pi}$ to $+{H_1\over 2\pi}$. Inside the bubble the field will be almost homogeneous; the inhomogeneity $\delta \phi \sim -{H_1\over 2\pi}$ will be concentrated only in a small vicinity near the bubble wall. Finally we should verify that this scenario leads to bubbles which are symmetric enough, see eq. (\ref{o17}). Fortunately, here we do not have any problems. One can easily check that for our model with $m \sim 10^{13}$ GeV and $\tilde M \sim \lambda^{-1/4} M > 10^{17} GeV$ the condition (\ref{o17}) can be satisfied even for not very small values of the coupling constant $\lambda$. The arguments presented above should be confirmed by a more detailed investigation of the vacuum structure inside the expanding bubble in our scenario. If, as we hope, the result of the investigation will be positive, we will have an extremely simple model of an open inflationary universe. In the meantime, it would be nice to have a model where we do not have any problems at all with synchronization and with large fluctuations on the scalar field in the false vacuum. We will consider such a model in the next section. \subsection{\label{Hybrid} Hybrid inflation and natural inflation with $\Omega < 1$} The model to be discussed below is a version of the hybrid inflation scenario \cite{Hybrid}, which is a slight generalization (and a simplification) of our previous model (\ref{3}): \begin{equation}\label{4a} V(\phi,\sigma) = {g^2\over 2}\phi^2\sigma^2 + V(\sigma) \ . \end{equation} We eliminated the massive term of the field $\phi$ and added explicitly the interaction ${g^2\over 2}\phi^2\sigma^2$, which, as we have mentioned already, can be useful (though not necessary) for stabilization of the state $\sigma = 0$ at large $\phi$. Note that in this model the line $\sigma = 0$ is a flat direction in the ($\phi,\sigma$) plane. At large $\phi$ the only minimum of the effective potential with respect to $\sigma$ is at the line $\sigma = 0$. To give a particular example, one can take $V(\sigma) = {M^2\over 2} \sigma^2 -{\alpha M } \sigma^3 + {\lambda\over 4}\sigma^4 +V_0$. Here $V_0$ is a constant which is added to ensure that $V(\phi,\sigma) = 0$ at the absolute minimum of $V(\phi,\sigma)$. In this case the minimum of the potential $V(\phi,\sigma)$ at $\sigma \not = 0$ is deeper than the minimum at $\sigma = 0$ only for $\phi < \phi_c$, where $\phi_c = {M\over g}\sqrt{{2\alpha^2\over \lambda} -1}$. This minimum for $\phi = \phi_c$ appears at $\sigma = \sigma_c = {2\alpha M\over \lambda}$. The bubble formation becomes possible only for $\phi < \phi_c$. After the tunneling the field $\phi$ acquires an effective mass $m = g\sigma$ and begins to move towards $\phi = 0$, which provides the mechanism for the second stage of inflation inside the bubble. In this scenario evolution of the scalar field $\phi$ is exactly synchronized with the evolution of the field $\sigma$, and the Universe inside the bubble appears to be open. Effective mass of the field $\phi$ at the minimum of $V(\phi,\sigma)$ with $\phi = \phi_c$, $\sigma = \sigma_c = {2\alpha M\over \lambda}$ is $m = g\sigma_c = {2g\alpha M\over \lambda}$. With a decrease of the field $\phi$ its effective mass at the minimum of $V(\phi,\sigma)$ will grow, but not significantly. For simplicity, we will consider the case $\lambda = \alpha^2$. In this case it can be shown that $V(0) = 2.77\, {M^4\over \lambda}$, and the Hubble constant before the phase transition is given by $4.8\, {M^2\over \sqrt \lambda M_{\rm P}}$. One should check what is necessary to avoid too large density perturbations (\ref{o17}). However, one should take into account that the mass $M$ in (\ref{o17}) corresponds to the curvature of the effective potential near $\phi = \phi_c$ rather than at $\phi = 0$. In our case this implies that one should use $\sqrt 2 M$ instead of $M$ in this equation. Then one obtains the following constraint on the mass $M$: \ $M\sqrt \mu {\ \lower-1.2pt\vbox{\hbox{\rlap{$<$}\lower5pt\vbox{\hbox{$\sim$}}}}\ } 2\times10^{15}$ GeV. Note that the thin wall approximation (requiring $\mu \ll 1$) breaks down far away from $\phi = \phi_c$. Therefore in general eq. (\ref{o17}) should be somewhat improved. However for $\phi \approx \phi_c$ it works quite well. To be on a safe side, we will take $M = 5\times 10^{14}$ GeV. Other parameters may vary; one may consider, e.g., the theory with $g \sim 10^{-5}$, which gives $\phi_c = {M\over g} \sim 5\times 10^{19}\ \mbox{GeV} \sim 4M_{\rm P}$. The effective mass $m$ after the phase transition is equal to ${2gM\over \sqrt \lambda}$ at $\phi = \phi_c$, and then it grows by only $25\%$ when the field $\phi$ changes all the way down from $\phi_c$ to $\phi = 0$. As we already mentioned, in order to obtain the proper amplitude of density perturbations produced by inflation inside the bubble one should have $m \sim 10^{13}$ GeV. This corresponds to $\lambda = \alpha^2 = 10^{-6}$. The bubble formation becomes possible only for $\phi < \phi_c$. If it happens in the interval $4M_{\rm P} > \phi > 3 M_{\rm P}$, we obtain a flat Universe. If it happens at $\phi < 3M_{\rm P}$, we obtain an open Universe. Depending on the initial value of the field $\phi$, we can obtain all possible values of $\Omega$, from $\Omega = 1$ to $\Omega = 0$. The value of the Hubble constant at the minimum with $\sigma \not = 0$ at $\phi = 3M_{\rm P}$ in our model does not differ much from the value of the Hubble constant before the bubble formation. Therefore we do not expect any specific problems with the large scale density perturbations in this model. Note also that the probability of tunneling at large $\phi$ is very small since the depth of the minimum at $\phi \sim \phi_c$, $\sigma \sim \sigma_c$ does not differ much from the depth of the minimum at $\sigma = 0$, and there is no tunneling at all for $\phi > \phi_c$. Therefore the number of flat Universes produced by this mechanism will be strongly suppressed as compared with the number of open Universes, the degree of this suppression being very sensitive to the value of $\phi_c$. Meanwhile, life of our type is impossible in empty Universes with $\Omega \ll 1$. This may provide us with a tentative explanation of the small value of $\Omega$ in the context of our model. Another model of inflation with $\Omega < 1$ is the based on a certain modification of the ``natural inflation'' scenario \cite{Natural}. The main idea is to take the effective potential of the ``natural inflation'' model, which looks like a tilted Mexican hat, and make a deep hole it its center at $\phi = 0$ \cite{Arthur}. In the beginning inflation occurs near $\phi = 0$, but then the bubbles with $\phi \not = 0$ appear. Depending on the phase of the complex scalar field $\phi$ inside the bubble, the next stage of inflation, which occurs just as in the old version of the ``natural inflation'' scenario, leads to formation of the Universes with all possible values of $\Omega$. A detailed discussion of this scenario can be found in \cite{Arthur}; we will not repeat it here. What is most important for us is that there exist several rather simple models of an open inflationary universe. Inflationary models with $\Omega = 1$ admittedly are somewhat simpler. Therefore we still hope that several years later we will know that our Universe is flat, which will be a strong experimental evidence in favor of inflationary cosmology in its simplest form. However, if observational data will show, beyond any reasonable doubt, that $\Omega \not = 1$, it will not imply that inflationary theory is wrong. Indeed, now we know that there is a large class of internally consistent cosmological models which may describe creation of large homogeneous Universes with all possible values of $\Omega$, and so far all of these models are based on inflationary cosmology. \ \section { Reheating after inflation} The theory of reheating of the Universe after inflation is the most important application of the quantum theory of particle creation, since almost all matter constituting the Universe at the subsequent radiation-dominated stage was created during this process \cite{MyBook}. At the stage of inflation all energy was concentrated in a classical slowly moving inflaton field $\phi$. Soon after the end of inflation this field began to oscillate near the minimum of its effective potential. Gradually it produced many elementary particles, they interacted with each other and came to a state of thermal equilibrium with some temperature $T_r$, which was called the reheating temperature. An elementary theory of reheating was first developed in \cite{DolgLinde} for the new inflationary scenario. Independently a theory of reheating in the $R^2$ inflation was constructed in~\cite{st81}. Various aspects of this theory were further elaborated by many authors, see e.g. \cite{Dolg,Brand}. Still, a general scenario of reheating was absent. In particular, reheating in the chaotic inflation theory remained almost unexplored. The present section contains results obtained recently in our work with Kofman and Starobinsky \cite{KLSREH}. We have found that the process of reheating typically consists of three different stages. At the first stage, which cannot be described by the elementary theory of reheating, the classical coherently oscillating inflaton field $\phi$ decays into massive bosons (in particular, into $\phi$-particles) due to parametric resonance. In many models the resonance is very broad, and the process occurs extremely rapidly (explosively). Because of the Pauli exclusion principle, there is no explosive creation of fermions. To distinguish this stage from the stage of particle decay and thermalization, we will call it {\it pre-heating}. Bosons produced at that stage are far away from thermal equilibrium and typically have enormously large occupation numbers. The second stage is the decay of previously produced particles. This stage typically can be described by methods developed in \cite{DolgLinde}. However, these methods should be applied not to the decay of the original homogeneous inflaton field, but to the decay of particles and fields produced at the stage of explosive reheating. This considerably changes many features of the process, including the final value of the reheating temperature. The third stage is the stage of thermalization, which can be described by standard methods, see e.g. \cite{MyBook,DolgLinde}; we will not consider it here. Sometimes this stage may occur simultaneously with the second one. In our investigation we have used the formalism of the time-dependent Bogoliubov transformations to find the density of created particles, $n_{\vec k}(t)$. A detailed description of this theory will be given in \cite{REH}; here we will outline our main conclusions using a simple semiclassical approach. We will consider a simple chaotic inflation scenario describing the classical inflaton scalar field $\phi$ with the effective potential $V(\phi) = \pm {1\over2} m_\phi^2 \phi^2+{\lambda\over 4}\phi^4$. Minus sign corresponds to spontaneous symmetry breaking $\phi \to \phi +\sigma$ with generation of a classical scalar field $\sigma = {m_\phi \over\sqrt\lambda}$. The field $\phi$ after inflation may decay into bosons $\chi$ and fermions $\psi$ due to the interaction terms $- { 1\over2} g^2 \phi^2 \chi^2$ and $- h \bar \psi \psi \phi$. Here $\lambda$, $ g$ and $h$ are small coupling constants. In case of spontaneous symmetry breaking, the term $- { 1\over2} g^2 \phi^2 \chi^2$ gives rise to the term $- g^2 \sigma\phi \chi^2$. We will assume for simplicity that the bare masses of the fields $\chi$ and $\psi$ are very small, so that one can write $ m_\chi (\phi) = g \phi$, $m_{\psi}(\phi) = |h\phi|$. Let us briefly recall the elementary theory of reheating \cite{MyBook}. At $\phi > M_{\rm P}$, we have a stage of inflation. This stage is supported by the friction-like term $3H\dot\phi$ in the equation of motion for the scalar field. Here $H\equiv \dot a/a$ is the Hubble parameter, $a(t)$ is the scale factor of the Universe. However, with a decrease of the field $\phi$ this term becomes less and less important, and inflation ends at $\phi {\ \lower-1.2pt\vbox{\hbox{\rlap{$<$}\lower5pt\vbox{\hbox{$\sim$}}}}\ }M_{\rm P}/2$. After that the field $\phi$ begins oscillating near the minimum of $V(\phi)$. The amplitude of the oscillations gradually decreases because of expansion of the Universe, and also because of the energy transfer to particles created by the oscillating field. Elementary theory of reheating is based on the assumption that the classical oscillating scalar field $\phi (t)$ can be represented as a collection of scalar particles at rest. Then the rate of decrease of the energy of oscillations coincides with the decay rate of $\phi$-particles. The rates of the processes $\phi \to \chi\chi$ and $\phi \to \psi\psi$ (for $m_\phi \gg 2m_\chi, 2m_\psi$) are given by \begin{equation}\label{7} \Gamma ( \phi \to \chi \chi) = { g^4 \sigma^2\over 8 \pi m_{\phi}}\ , \ \ \ \ \ \Gamma( \phi \to \psi \psi ) = { h^2 m_{\phi}\over 8 \pi}\ . \end{equation} Reheating completes when the rate of expansion of the Universe given by the Hubble constant $H=\sqrt{8\pi \rho\over 3 M^2_{\rm P}} \sim t^{-1}$ becomes smaller than the total decay rate $\Gamma = \Gamma (\phi \to \chi \chi) + \Gamma (\phi \to \psi \psi )$. The reheating temperature can be estimated by $T_r \simeq 0.1\, \sqrt{\Gamma M_{\rm P}}$\,. It is interesting to note that in accordance with the elementary theory of reheating the amplitude squared of the oscillating scalar field decays exponentially, as $e^{-\Gamma t}$. Phenomenologically, this can be described by adding the term $\Gamma\dot\phi$ to the equation of motion of the scalar field. Unfortunately, many authors took this prescription too seriously and investigated the possibility that the term $\Gamma\dot\phi$, just like the term $3H\dot\phi$, can support inflation. We should emphasize \cite{KLSREH}, that adding the term $\Gamma\dot\phi$ to the equation of motion is justified only at the stage of oscillations (i.e. after the end of inflation), and only for the description of the {\it amplitude of oscillations} of the scalar field, rather than for the description of the scalar field itself. Moreover, even at the stage of oscillations this description becomes incorrect as soon as the resonance effects become important. As we already mentioned, elementary theory of reheating can provide a qualitatively correct description of particle decay at the last stages of reheating. Moreover, this theory is always applicable if the inflaton field can decay into fermions only, with a small coupling constant $h^2 \ll m_{\phi}/M_{\rm P}$. However, typically this theory is inapplicable to the description of the first stages of reheating, which makes the whole process quite different. In what follows we will develop the theory of the first stages of reheating. We will begin with the theory of a massive scalar field $\phi$ decaying into particles $\chi$, then we consider the theory ${\lambda\over 4} \phi^4$, and finally we will discuss reheating in the theories with spontaneous symmetry breaking. We begin with the investigation of the simplest inflationary model with the effective potential ${m^2_\phi\over 2}\phi^2$. Suppose that this field only interacts with a light scalar field $\chi$ ($m_{\chi} \ll m_{\phi}$) due to the term $-{ 1\over2} g^2 \phi^2 \chi^2$. The equation for quantum fluctuations of the field $\chi$ with the physical momentum $\vec k/a(t)$ has the following form: \begin{equation}\label{M} \ddot \chi_k + 3H \dot \chi_k + \left({k^2\over a^2(t)} + g^2 \Phi^2\, \sin^2(m_{\phi}t) \right) \chi_k = 0 \ , \end{equation} where $k = \sqrt {\vec k^2}$, and $\Phi$ stands for the amplitude of oscillations of the field $\phi$. As we shall see, the main contribution to $\chi$-particle production is given by excitations of the field $\chi$ with $k/a \gg m_\phi$, which is much greater than $H$ at the stage of oscillations. Therefore, in the first approximation we may neglect the expansion of the Universe, taking $a(t)$ as a constant and omitting the term $3H \dot \chi_k$ in (\ref{M}). Then the equation (\ref{M}) describes an oscillator with a variable frequency $\Omega_k^2(t)= k^2a^{-2} + g^2\Phi^2\, \sin^2(m_{\phi}t) $. Particle production occurs due to a nonadiabatic change of this frequency. Equation (\ref{M}) can be reduced to the well-known Mathieu equation: \begin{equation}\label{M1} \chi_k'' + \left(A(k) - 2q \cos 2z \right) \chi_k = 0 \ , \end{equation} where $A(k) = {k^2 \over m_\phi^2 a^2}+2q$, $q = {g^2\Phi^2\over 4m_\phi^2} $, $z = m_{\phi}t$, prime denotes differentiation with respect to $z$. An important property of solutions of the equation (\ref{M1}) is the existence of an exponential instability $\chi_k \propto \exp (\mu_k^{(n)}z)$ within the set of resonance bands of frequencies $\Delta k^{(n)}$ labeled by an integer index $n$. This instability corresponds to exponential growth of occupation numbers of quantum fluctuations $n_{\vec k}(t) \propto \exp (2\mu_k^{(n)} m_{\phi} t)$ that may be interpreted as particle production. As one can show, near the line $A = 2q$ there are regions in the first, the second and the higher instability bands where the unstable modes grow extremely rapidly, with $\mu_k \sim 0.2$. We will show analytically in~\cite{REH} that for $q \gg 1$ typically $\mu_k \sim {\ln 3\over 2\pi} \approx 0.175$ in the instability bands along the line $A = 2q$, but its maximal value is ${\ln(1+\sqrt{2}) \over \pi} \approx 0.28$. Creation of particles in the regime of a broad resonance ($q > 1$) with $2\pi \mu_k = O(1)$ is very different from that in the usually considered case of a narrow resonance ($ q \ll 1$), where $2\pi \mu_k \ll 1$. In particular, it proceeds during a tiny part of each oscillation of the field $\phi$ when $1-\cos z \sim q^{-1}$ and the induced effective mass of the field $\chi$ (which is determined by the condition $m^2_{\chi}= g^2\Phi^2/2$) is less than $m_{\phi}$. As a result, the number of particles grows exponentially within just a few oscillations of the field $\phi$. This leads to an extremely rapid (explosive) decay of the classical scalar field $\phi$. This regime occurs only if $q {\ \lower-1.2pt\vbox{\hbox{\rlap{$>$}\lower5pt\vbox{\hbox{$\sim$}}}}\ } \pi^{-1}$, i.e. for $g\Phi {\ \lower-1.2pt\vbox{\hbox{\rlap{$>$}\lower5pt\vbox{\hbox{$\sim$}}}}\ } m_\phi$, so that $m_\phi \ll gM_{\rm P}$ is the necessary condition for it. One can show that a typical energy $E$ of a particle produced at this stage is determined by equation $A-2q \sim \sqrt{q}$, and is given by $E \sim \sqrt{g m_\phi M_{\rm P}}$ \cite{REH}. Creation of $\chi$-particles leads to the two main effects: transfer of the energy from the homogeneous field $\phi (t)$ to these particles and generation of the contribution to the effective mass of the $\phi$ field: $m^2_{\phi ,eff}=m^2_{\phi}+g^2\langle\chi^2 \rangle_{ren}$. The last term in the latter expression quickly becomes larger than $m^2_{\phi}$. One should take both these effects into account when calculating backreaction of created particles on the process. As a result, the stage of the broad resonance creation ends up within the short time $t\sim m_{\phi}^{-1} \ln (m_{\phi}/g^5M_{\rm P})$, when $\Phi^2 \sim \langle\chi^2\rangle$ and $q = {g^2\Phi^2\over 4m_{\phi ,eff}^2}$ becomes smaller than $1$. At this time the energy density of produced particles $\sim E^2 \langle\chi^2\rangle \sim g m_\phi M_{\rm P} \Phi^2$ is of the same order as the original energy density $\sim {m_\phi^2} M_{\rm P}^2$ of the scalar field $\phi$ at the end of inflation. This gives the amplitude of oscillations at the end of the stage of the broad resonance particle creation: $\Phi^2 \sim \langle\chi^2\rangle \sim g^{-1} m_\phi M_{\rm P} \ll M_{\rm P}^2$. Since $E\gg m_{\phi}$, the effective equation of state of the whole system becomes $p\approx \rho /3$. Thus, explosive creation practically eliminates a prolonged intermediate matter-dominated stage after the end of inflation which was thought to be characteristic to many inflationary models. However, this does not mean that the process of reheating has been completed. Instead of $\chi$-particles in the thermal equilibrium with a typical energy $E \sim T \sim (mM_{\rm P})^{1/2}$, one has particles with a much smaller energy $\sim (g m_\phi M_{\rm P})^{1/2}$, but with extremely large mean occupation numbers $n_k \sim g^{-2} \gg 1$. After that the Universe expands as $a(t)\propto \sqrt t$, and the scalar field $\phi$ continues its decay in the regime of the narrow resonance creation $q\approx {\Phi^2\over 4 \langle\chi^2\rangle} \ll 1$. As a result, $\phi$ decreases rather slowly, $\phi \propto t^{-3/4}$. This regime is very important because it makes the energy of the $\phi$ field much smaller than that of the $\chi$-particles. One can show that the decay finally stops when the amplitude of oscillations $\Phi$ becomes smaller than $g^{-1} m_\phi$ \cite{REH}. This happens at the moment $t\sim m_{\phi}^{-1} (gM_{\rm P}/m_{\phi})^{1/3}$ (in the case $m < g^7 M_{\rm P}$ decay ends somewhat later, in the perturbative regime). The physical reason why the decay stops is rather general: decay of the particles $\phi$ in our model occurs due to its interaction with another $\phi$-particle (interaction term is quadratic in $\phi$ and in $\chi$). When the field $\phi$ (or the number of $\phi$-particles) becomes small, this process is inefficient. The scalar field can decay completely only if a single scalar $\phi$-particle can decay into other particles, due to the processes $\phi \to \chi \chi$ or $\phi \to \psi \psi$, see eq. (\ref{7}). If there is no spontaneous symmetry breaking and no interactions with fermions in our model, such processes are impossible. At later stages the energy of oscillations of the inflaton field decreases as $a^{-3}(t)$, i.e. more slowly than the decrease of energy of hot ultrarelativistic matter $\propto a^{-4}(t)$. Therefore, the relative contribution of the field $\phi(t)$ to the total energy density of the Universe rapidly grows. This gives rise to an unexpected possibility that the inflaton field by itself, or other scalar fields can be cold dark matter candidates, {\it even if they strongly interact with each other}. However, this possibility requires a certain degree of fine tuning; a more immediate application of our result is that it allows one to rule out a wide class of inflationary models which do not contain interaction terms of the type of $g^2\sigma\phi\chi^2$ or $h\phi\bar\psi\psi$. So far we have not considered the term ${\lambda \over 4} \phi^4$ in the effective potential. Meanwhile this term leads to production of $\phi$-particles, which in some cases appears to be the leading effect. Let us study the $\phi$-particle production in the theory with $V(\phi) = {m^2_{\phi}\over 2} \phi^2 + {\lambda\over 4}\phi^4$ with $m^2_{\phi} \ll \lambda M_{\rm P}^2$. In this case the effective potential of the field $\phi$ soon after the end of inflation at $\phi \sim M_{\rm P}$ is dominated by the term ${\lambda\over 4} \phi^4$. Oscillations of the field $\phi$ in this theory are not sinusoidal, they are given by elliptic functions, but with a good accuracy one can write $\phi(t) \sim \Phi \sin (c\sqrt \lambda \int \Phi dt)$, where $c={\Gamma^2(3/4)\over \sqrt \pi} \approx 0.85$. the Universe at that time expands as at the radiation-dominated stage: $a(t)\propto \sqrt t$. If one neglects the feedback of created $\phi$-particles on the homogeneous field $\phi (t)$, then its amplitude $\Phi (t) \propto a^{-1}(t)$, so that $a\Phi =const$. Using a conformal time $\eta$, exact equation for quantum fluctuations $\delta \phi$ of the field $\phi$ can be reduced to the Lame equation. The results remain essentially the same if we use an approximate equation \begin{equation}\label{lam1} {d^2(\delta\phi_k)\over d\eta^2} + {\Bigl[{k^2} + 3\lambda a^2\Phi^2\, \sin^2 (c\sqrt\lambda a\Phi \eta)\Bigr]} \delta\phi_k = 0,~~\eta =\int {dt\over a(t)}={2t\over a(t)}\, , \end{equation} which leads to the Mathieu equation with $A = {k^2\over c^2\lambda a^2\Phi^2} + {3\over 2c^2} \approx {k^2\over c^2\lambda a^2\Phi^2} + 2.08$, and $q = {3\over 4c^2} \approx 1.04$. Looking at the instability chart, we see that the resonance occurs in the second band, for $k^2 \sim 3\lambda a^2\Phi^2$. The maximal value of the coefficient $\mu_k$ in this band for $q \sim 1$ approximately equals to $0.07$. As long as the backreaction of created particles is small, expansion of the Universe does not shift fluctuations away from the resonance band, and the number of produced particles grows as $\exp (2c\mu_k\sqrt\lambda a\Phi \eta) \sim \exp ({\sqrt\lambda\Phi t\over 5})$. After the time interval $\sim M_{\rm P}^{-1}\lambda^{-1/2}|\ln \lambda|$, backreaction of created particles becomes significant. The growth of the fluctuations $\langle\phi^2\rangle$ gives rise to a contribution $3\lambda \langle\phi^2\rangle$ to the effective mass squared of the field $\phi$, both in the equation for $\phi (t)$ and in Eq. (\ref{lam1}) for inhomogeneous modes. The stage of explosive reheating ends when $\langle\phi^2\rangle$ becomes greater than $\Phi^2$. After that, $\Phi^2 \ll \langle\phi^2\rangle$ and the effective frequency of oscillations is determined by the term $\sqrt{3\lambda \langle\phi^2\rangle}$. The corresponding process is described by Eq. (\ref{M1}) with $A(k) = 1 + 2q + {k^2 \over 3\lambda a^2\langle\phi^2\rangle}$, $q = {\Phi^2\over 4 \langle\phi^2\rangle}$. In this regime $q \ll 1$, and particle creation occurs in the narrow resonance regime in the second band with $A \approx 4$. Decay of the field in this regime is extremely slow: one can show \cite{REH} that the amplitude $\Phi$ decreases only by a factor $t^{1/12}$ faster that it would decrease without any decay, due to the expansion of the Universe only, i.e., $\Phi \propto t^{-7/12}$. Reheating stops altogether when the presence of non-zero mass $m_{\phi}$ though still small as compared to $\sqrt{3\lambda \langle\phi^2\rangle}$ appears enough for the expansion of the Universe to drive a mode away from the narrow resonance. It happens when the amplitude $\Phi$ drops up to a value $\sim m_{\phi}/\sqrt \lambda$. In addition to this process, the field $\phi$ may decay to $\chi$-particles. This is the leading process for $g^2\gg \lambda$. The equation for $\chi_k$ quanta has the same form as eq. (\ref{lam1}) with the obvious change $\lambda \to g^2/3$. Initially parametric resonance is broad. The values of the parameter $\mu_k$ along the line $A = 2q$ do not change monotonically, but typically for $q \gg 1$ they are 3 to 4 times greater than the parameter $\mu_k$ for the decay of the field $\phi$ into its own quanta. Therefore, this pre-heating process is very efficient. It ends at the moment $t\sim M_{\rm P}^{-1}\lambda^{-1/2} \ln (\lambda /g^{10})$ when $\Phi^2 \sim \langle \chi^2 \rangle \sim g^{-1}\sqrt \lambda M_{\rm P}^2$. The typical energy of created $\chi$-particles is $E \sim (g^2\lambda)^{1/4}M_{\rm P}$. The following evolution is essentially the same as that described above for the case of a massive scalar field decaying into $\chi$-particles. Finally, let us consider the case with symmetry breaking. In the beginning, when the amplitude of oscillations is much greater than $\sigma$, the theory of decay of the inflaton field is the same as in the case considered above. The most important part of pre-heating occurs at this stage. When the amplitude of the oscillations becomes smaller than $m_\phi/\sqrt\lambda$ and the field begins oscillating near the minimum of the effective potential at $\phi = \sigma$, particle production due to the narrow parametric resonance typically becomes very weak. The main reason for this is related to the backreaction of particles created at the preceding stage of pre-heating on the rate of expansion of the Universe and on the shape of the effective potential \cite{REH}. A rather interesting effect which makes investigation of this regime especially complicated is a temporary (non-thermal) symmetry restoration which occurs because of the interaction of the field $\phi$ with its fluctuations $<\phi^2>$. Importance of spontaneous symmetry breaking for the theory of reheating should not be underestimated, since it gives rise to the interaction term $g^2\sigma\phi\chi^2$ which is linear in $\phi$. Such terms are necessary for a complete decay of the inflaton field in accordance with the perturbation theory (\ref{7}). In this section we presented the new theory of reheating developed in \cite{KLSREH}, where we performed an investigation of reheating with an account of expansion of the Universe and of the backreaction of created particles, both in the broad resonance regime and in the narrow resonance case. As a result of this investigation, we obtained equations for the power-law decrease of the amplitude of an oscillating scalar field with an account taken of all of these effects. During the last year there appeared many other papers on the theory of reheating \cite{Shtanov}--\cite{Kaiser}, which made the physical picture of reheating even more clear. Unfortunately, it is not easy to compare the results obtained in \cite{Shtanov}--\cite{Kaiser} with the results of our work \cite{KLSREH}. For example, a very thorough investigation of reheating in the narrow resonance regime without a complete account of backreaction was performed in \cite{Shtanov,Yoshimura,Kaiser}, and their results in this approximation agree with the corresponding results of \cite{KLSREH}. However, as we have seen, at the first, most efficient stages of reheating the resonance is broad, and when it becomes narrow a complete account of backreaction becomes necessary \cite{KLSREH}. Backreaction was studied in a very detailed way in ref.~\cite{Boyan}, but their investigation was performed neglecting expansion of the Universe, which was an important part of our work. That is why in this review we concentrated on the results obtained in \cite{KLSREH}. However, to obtain a complete theory of reheating a much more detailed investigation will be necessary, and in this respect many of the results obtained in \cite{Shtanov}--\cite{Kaiser} should be very useful. We should emphasize that the stage of parametric resonance is just the first stage of the process. If one naively takes the energy density at the end of explosive reheating and assumes that this energy density instantaneously transfers to heat, one may overestimate the reheating temperature by many orders of magnitude. Indeed, after the stage of explosive reheating the bose-particles created at this stage have enormously large occupation numbers, and they should further decay into the usual elementary particles. This may take a lot of time, during which the energy density of the Universe may decrease dramatically. To find the reheating temperature one should investigate the subsequent decay of the particles created at the stage of explosive reheating. This decay can be described by the old perturbative methods developed in~\cite{DolgLinde}. Note, however, that now this theory should be applied not to the decay of the original large and homogeneous oscillating inflaton field, but to the decay of particles produced at the stage of pre-heating, as well as to the decay of small remnants of the classical inflaton field. This makes a lot of difference, since typically coupling constants of interaction of the inflaton field with matter are extremely small, whereas coupling constants involved in the decay of other bosons can be much greater. As a result, the reheating temperature can be much higher than the typical temperature $T_r {\ \lower-1.2pt\vbox{\hbox{\rlap{$<$}\lower5pt\vbox{\hbox{$\sim$}}}}\ } 10^9$ GeV which could be obtained neglecting the stage of parametric resonance \cite{REH}. In addition, one should make a careful study of the process of establishing of thermal equilibrium \cite{Boyan2}. On the other hand, such processes as baryon creation after inflation occur best of all outside the state of thermal equilibrium. Therefore, the stage of explosive reheating (pre-heating), which produces fields and particles outside of the state of thermal equilibrium, may play an extremely important role in the cosmological theory. Another consequence of the resonance effects is an almost instantaneous change of equation of state from the vacuum-like one to the equation of state of relativistic matter $p = \rho/3$. This may be important for investigation of the primordial black hole formation, which could appear from growing density perturbations if equation of state after inflation for a long time was $p = 0$. A rather nontrivial example of reheating appears in inflationary models based on supergravity, see e.g. \cite{Nan}--\cite{Bert}. The leading mode of the single-inflaton decay in such models often involves creation of a gravitino, which is a fermion. This does not necessarily mean that the first explosive stage cannot be realized in such models. Indeed, just as in the theory ${\lambda\over 4}\phi^4$, at the first stage the homogeneous classical oscillating inflaton field $\phi$ may decay into decoherent waves or particles of the same field $\phi$. However, this will be just a first stage of reheating, after which one should consider decay of the inflaton particles by the usual perturbative methods. In such a situation one does not expect any deviations of the reheating temperature from its value obtained by perturbative methods \cite{DolgLinde}, \cite{Nan}--\cite{Bert}. One should note also that in certain models the oscillations of the scalar field from the very beginning occur in the region where the conditions for the explosive reheating formulated in \cite{KLSREH} are not satisfied. Such a situation occurs, e.g., in ``natural inflation'' \cite{DolgF}, where the change of the effective mass of the inflaton field during its oscillations is relatively small, and the conditions of existence of narrow resonance in expanding Universe derived in \cite{KLSREH} are violated. Let us briefly summarize our results: 1. In many models where decay of the inflaton field can occur in the purely bosonic sector the first stages of reheating occur due to parametric resonance. This process (pre-heating) is extremely efficient even if the corresponding coupling constants are very small. However, there is no explosive reheating in the models where decay of the inflaton field is necessarily accompanied by fermion production. 2. The stage of explosive reheating due to a broad resonance typically is very short. Later the resonance becomes narrow, and finally the stage of pre-heating finishes altogether. Interactions of particles produced at this stage, their decay into other particles and subsequent thermalization typically require much more time that the stage of pre-heating, since these processes are suppressed by the small values of coupling constants. 3. The last stages of reheating typically can be described by the elementary theory of reheating \cite{DolgLinde}. However, this theory should be applied not to the original inflaton field, but to the products of its decay formed at the stage of explosive reheating. In some models it changes the final value of the reheating temperature. 4. Existence of the intermediate stage between the end of explosive reheating and the beginning of thermal equilibrium may have important implications for the theory of baryogenesis. 5. Reheating never completes in the theories where a single $\phi$-particle cannot decay into other particles. This implies that reheating completes only if the theory contains interaction terms like $\phi\sigma\chi^2$ of $\phi\bar\psi\psi$. In most cases the theories where reheating never completes contradict observational data. On the other hand, this result suggests an interesting possibility that the classical scalar fields (maybe even the inflaton field itself) may be responsible for the dark matter of the Universe even if they strongly interact with other matter fields. \section{Conclusions} Inflationary theory is already more than 15 years old, and its main principles seem to be well understood. Nevertheless, it is young enough to bring us many new surprises. Originally we expected that inflation was a short intermediate stage after the hot big bang. Now it seems that the standard big bang theory is only a part of inflationary cosmology which describes local (but not global) properties of the self-reproducing inflationary universe. Even though each part of the Universe expands (or collapses), the Universe as a whole may be stationary. One of the main purposes of inflationary cosmology was to solve the primordial monopole problem by expanding the distance between the monopoles. Recently we learned that the monopoles themselves may expand exponentially and become as large as a universe \cite{LL}. On the other hand, we learned that an infinitely large open inflationary universe may fit into an interior of a single bubble of a finite size produced during the false vacuum decay. This demonstrated that even though $\Omega = 1$ remains one of the rather robust predictions of inflationary cosmology, it will be impossible to kill inflation by proving that our universe is open. The process of creation of matter after inflation also happened to be extremely interesting and complicated, involving investigation of nonperturbative resonance effects in an expanding universe. Rapid development of the inflationary theory is a very good sign indicating that we are moving fast towards a complete cosmological theory -- assuming, as we all hope, that we have chosen the right direction. \newpage
\section{Introduction} Geometric approaches to quantum mechanics have been studied by various groups ever since the foundation of quantum mechanics was laid down. The prime aim of such approaches is to render quantum mechanics applicable to more general settings, not just to Euclidean space as originally done. However, it is by now well recognized that quantization is generally difficult to carry out unless the setting is fairly simple. A system whose classical configuration space $Q$ is a homogeneous space given by a coset $G/H$ falls into this simple category. An important lesson learned when quantizing on homogeneous spaces is that there are actually (infinitely) many {\it inequivalent quantizations} allowed~\cite{Mackey,Doebner,Isham}. In other words, there exist unitarily inequivalent Hilbert spaces where physical properties, such as their energy spectra, may differ from each other. These inequivalent quantizations are classified according to the induced representation~\cite{Mackey} which is used for the quantization. \par Interest in the inequivalent quantizations has been renewed recently after Landsman and Linden examined the physical implications of the quantizations and found that a special type of gauge field is induced on homogeneous spaces~\cite{Landsman,LL,Ohnuki,McMullan,Robson}. The gauge fields are a (topological) solution of the Yang-Mills equation on the spaces, called the {\it canonical connection} (or $H$-{\it connection}). However, the previous arguments leading to the gauge fields are algebraic and abstract, and there is no intuitive account of this rather mysterious appearance of gauge fields. It would be therefore desirable to develop a path integral account, which normally admits a more intuitive understanding based on the geometry of the configuration space. In this note we wish to take a step in this direction --- we shall show that, the path integral on a homogeneous space carries the canonical connection as a gauge field, if we adopt the guiding principle that the path integral be constructed first on the group manifold $G$ and then projected down to the homogeneous space $G/H$. This \lq first lift and then project' principle may be arguable, but it is certainly true that the case $Q = S^1$, where the path integral is known to reproduce the inequivalent quantizations correctly~\cite{Schulman}, relies on this principle. We shall not dwell on this issue until the end of the paper where a possible explanation is given. We here mention that similar induced gauge fields appear in various other contexts as well; {\it e.g.}, in the context of Berry's (geometric) phase in quantum mechanics~\cite{Berry,Simon,Wilczek,Levay} or in the kinematics of molecules and deformable bodies~\cite{Guichardet,Iwai,Shapere,Montgomery}. Moreover, induced gauge fields play an important role in high energy physics too; {\it e.g.}, in the so-called hidden local symmetry of nonlinear sigma models~\cite{Bando} and in the search for a possible origin of dynamical gauge bosons~\cite{Kikkawa,Tamura}. We hope that the path integral account given in this paper may shed some light on the machinery for those phenomena in general. \par The plan of this paper is as follows. First we review quantum mechanics on $Q = S^1 $ to see how the gauge field is induced. Motivated by this simple example, we then generalize the construction of the path integral to the case of homogeneous spaces $ Q = G/H $ following the above guiding principle. We shall find that the canonical connection does appear in the path integral in the form expected, once the induced representation is incorporated in the path integral scheme. Finally we will argue a possible generalization to inhomogeneous spaces, together with a restriction that may underlie the guiding principle we adopted. \section{Covering the path integral} Let us begin by reviewing the path integral on a circle $ S^1 $ \cite{Schulman}. (A further discussion can be found in refs.\cite{LL,S1}.) We first regard $ S^1 $ as the coset $S^1 \cong \mbox{\boldmath $ R $} / 2 \pi \mbox{\boldmath $ Z $} $ by identifying the point $ x $ of $ \mbox{\boldmath $ R $} $ with other points $ x + 2 \pi n$ for $ n \in \mbox{\boldmath $ Z $}$. This identification defines a covering map $ \pi : \mbox{\boldmath $ R $} \to S^1 $. Our idea is then to construct the path integral on $S^1$ from the path integral on $\mbox{\boldmath $ R $}$ with the above identification in mind. \par Let $ K_{R} ( x', x ; t ) = \langle x' | e^{ -i H t } | x \rangle $ be a propagator on $ \mbox{\boldmath $ R $} $ which is invariant under the translation by $2\pi$, \begin{equation} K_{R} ( x'+ 2 \pi , x + 2 \pi ; t ) = K_{R} ( x' , x ; t ). \label{0} \end{equation} On account of the identification of points $ x' + 2 \pi n $ with $ x' $, summation over $ n $ may lead to a propagator on $ S^1 $; \begin{equation} K_{S^1} ( x', x ; t ) = \sum_{ n = - \infty }^{ \infty } K_{R} ( x' + 2 \pi n , x ; t ), \label{1} \end{equation} where we interpret the integer $ n $ as the winding number of a path connecting two points $ x $ and $ x' $ along the circle $ S^1 $. Clearly, this expression admits an immediate generalization. In fact, we do not have an {\it a priori} physical reason to add up propagators for different winding numbers with the same weight, as long as the weight is a phase factor. Based on this observation Schulman~\cite{Schulman} proposed to insert a weight factor $ \omega_n $ with $ | \omega_n | = 1 $ to obtain a more general propagator \begin{equation} K_{S^1}^{\omega} ( x' , x ; t ) = \sum_{ n = - \infty }^{ \infty } \, \omega_n K_{R} ( x' + 2 \pi n , x ; t ). \label{2} \end{equation} The composition law of the propagator \begin{equation} \int_{0}^{2 \pi} dx' \, K_{S^1}^{\omega} ( x'', x' ; t' ) \, K_{S^1}^{\omega} ( x' , x ; t ) = K_{S^1}^{\omega} ( x'', x ; t + t' ) \label{3} \end{equation} is guaranteed if the weight satisfies\footnote{% The weight can actually be determined by requiring consistency against a shift of winding numbers~\cite{Schulman} but here we use the composition law for our later generalization.} \begin{equation} \omega_m \, \omega_n = \omega_{m+n}. \label{4} \end{equation} This implies that $ \omega : \pi_1 ( S^1 ) \to U(1) $ is a unitary representation of the first homotopy group $ \pi_1 ( S^1 ) \cong \mbox{\boldmath $ Z $} $ and hence given by $ \omega_n = e^{ i \alpha n } $ with a real parameter $ \alpha \in [ 0 , \, 2 \pi ) $. For each value of $ \alpha $, the propagator (\ref{2}) furnishes an inequivalent quantum theory on $ S^1 $. To see the physical meaning of $ \alpha $, we assume $ K_{R} $ to be of the standard form \begin{equation} K_{R} ( x' , x ; t ) = \int_x^{x'} [dx] \, \exp \left[ i \int dt \biggl\{ \frac{1}{2} \biggl( \frac{ d x }{ d t } \biggr)^2 - V ( x ) \biggr\} \right], \label{5} \end{equation} where $ V( x + 2 \pi ) = V( x ) $ in order to satisfy (\ref{0}). Putting $ A = \alpha / ( 2 \pi ) $, we find that the propagator (\ref{2}) can be rewritten as \begin{equation} K_{S^1}^{\omega} ( x' , x ; t ) = e^{-i\frac{\alpha}{2\pi}( x' - x )} \sum_{ n = - \infty }^{ \infty } \, \int_x^{ x' + 2 \pi n } [ dx ] \, \exp \left[ i \int dt \biggl\{ \frac{1}{2} \biggl( \frac{ d x }{ d t } \biggr)^2 - V ( x ) + A \, \frac{ d x }{ d t } \biggr\} \right]. \label{6} \end{equation} We therefore see that the insertion of the weight $ \omega_n = e^{ i \alpha n } $ just amounts to introduction of the minimal coupling with the vector potential $ A $. Being constant, the vector potential has vanishing curvature on $S^1$ but the flux penetrating the circle is finite. Hence, its physical consequence is analogous to that of the Aharonov-Bohm effect. \section{Lifting the path integral} What we have considered above is a covering $ \pi : \mbox{\boldmath $ R $} \to S^1 \cong \mbox{\boldmath $ R $} / 2 \pi \mbox{\boldmath $ Z $} $. A point $ x' $ in $ S^1 $ is lifted to points $ x' + 2 \pi n $ in $ \mbox{\boldmath $ R $} $, which are translated by the action of the group $ \mbox{\boldmath $ Z $} $. For each lifted point a propagator in $ \mbox{\boldmath $ R $} $ is defined, then we add them up with a weight factor given by the representation $ \omega : \mbox{\boldmath $ Z $} \to U(1) $ to obtain a propagator in $ S^1 $. Thus a path in $ S^1 $ is lifted up to $ \mbox{\boldmath $ R $} $ once, and then it is projected down to $ S^1 $ with a nontrivial weight multiplied, resulting in inequivalent quantizations and inducing a $ U(1) $ gauge field. In this section, we shall repeat the above construction of the path integral to a homogeneous space $G/H$, where $ G $ is a compact Lie group and $ H $ its closed subgroup. In order to set up a framework where a generalization of the covering $ \mbox{\boldmath $ R $} \to S^1 \cong \mbox{\boldmath $ R $} / 2 \pi \mbox{\boldmath $ Z $} $ can be realized for $Q \cong G/H$, we take the principal fiber bundle $ \pi : G \to G/H $ in which $ H $ acts on $ G $ from the right and $ G $ acts on $ G/H $ from the left. The difference from the former case is that $ H $ can be a continuous group or a nonabelian group in general, and hence the summation over the winding numbers $ \sum_n \, ( n \in \mbox{\boldmath $ Z $} ) $ will be replaced by the integration over the group $ \int_H \, d h $. For a nonabelian $ H $ its 1-dimensional representation is always trivial, but if we use higher dimensional nontrivial representations we will get inequivalent quantizations, as we shall see below. \par According to our guiding principle, we first lift our system from $Q$ to $ G $, and consider a propagator in $ G $ which is a map $ K_G : G \times G \times \mbox{\boldmath $ R $}^+ \to \mbox{\boldmath $ C $} $. The propagator we are interested in is one which is invariant under the $H$ action (as in (\ref{0})), \begin{equation} K_G ( g'h , g h ; t ) = K_G ( g' , g ; t )\ , \label{7} \end{equation} for arbitrary $ g, \, g' \in G $ and $ h \in H $. As before, we take the standard form for the propagator on $ G $, \begin{equation} K_G ( g' , g ; t ) = \int_g^{g'} [dg] \, \exp \left[ i \int dt \biggl\{ \frac{1}{2} \biggl| \biggl| \frac{ d g }{ d t } \biggr| \biggr|^2 - V ( g ) \biggr\} \right]\ . \label{8} \end{equation} Since the condition (\ref{7}) implies the invariance of the potential $ V( g h ) = V( g ) $, which corresponds to the periodicity $ V( x + 2 \pi ) = V( x ) $ in (\ref{5}), the potential $ V $ is actually a function of the homogeneous space $ V : Q \to \mbox{\boldmath $ R $} $. The norm $ || \cdot || $ used in (\ref{8}) is given by the invariant metric on $ G $, that is, $ || \dot{g} ||^2 = \mbox{Tr} ( g^{-1} \dot{g} )^2 $ where \lq $\mbox{Tr}$' is a matrix trace properly normalized in some irreducible representation. (The expression (\ref{8}) is rather symbolic; for a concrete expression, see \cite{LL}.) Now we define two unitary operators $ U_t $ and $ R_h $ acting on $ \psi \in L_2 ( G ) $ by \begin{eqnarray} && ( U_t \psi ) ( g' ) = \int_G dg \, K_G ( g' , g ; t ) \, \psi ( g )\ , \label{10} \\ && ( R_h \psi ) ( g ) = \psi ( g h )\ , \label{11} \end{eqnarray} for each $ t > 0 $ and $ h \in H $, where $ dg $ in (\ref{10}) is the normalized Haar measure of $ G $. Then, the invariance (\ref{7}) states that $ U_t \, R_h = R_h \, U_t $, and hence there exists a conserved quantity associated with this invariance. Consequently, the Hilbert space $ L_2(G) $ can be decomposed into the irreducible representations of $ H $. \par To implement the decomposition, let $ ( V_\chi , \rho_\chi ) $ be an irreducible unitary representation of $ H $, where $ V_\chi $ is a representation space labeled by $ \chi $. A function $ f : G \to V_\chi $ is called {\it $ \chi $-equivariant} if it satisfies $ f ( g h ) = \rho_\chi( h )^{-1} \, f ( g ) $. In other words, $ f $ is a section of the associated vector bundle $ E_\chi = G \times_\rho \, V_\chi $. The space of $ \chi $-equivariant functions is denoted by $ \Gamma^\chi $, which is equipped with the inner product \begin{equation} \langle f_1 , f_2 \rangle = \int_G dg \, \langle f_1 ( g ) , f_2 ( g ) \rangle, \label{12} \end{equation} where in the right-hand side $ \langle \cdot , \cdot \rangle $ denotes the inner product of the linear space $ V_\chi $. Consider then the operator $ I^{ ( \chi , j ) } : L_2(G) \to \Gamma^\chi $ defined by \begin{equation} ( I^{ ( \chi , j ) } \psi )^i ( g ) = \sqrt{ d_\chi } \int_H dh \, \rho_{\chi}^{ ij } ( h ) \, \psi ( g h )\ . \label{13} \end{equation} Here $ d_\chi = \dim V_\chi $, the indices $ i, j = 1, \cdots , d_\chi $ run over the components of $ V_\chi $, $ \rho_{\chi}^{ ij } ( h ) $ is a matrix element of an unitary representation $ \rho_{\chi} ( h ) $, and $ dh $ the normalized Haar measure of $ H $. This operator $ I^{ ( \chi , j ) } $ provides a partial isometry in the sense that $ I^{ ( \chi , j ) } $ is isometric on $ ( \ker I^{ ( \chi , j ) } )^{\perp} $ (for more detail on $ I^{ ( \chi , j ) } $, see \cite{LL}). The adjoint operator $ I^{ ( \chi , j ) \dagger } : \Gamma^\chi \to L_2(G) $ is defined by the relation $ \langle I^{ ( \chi , j ) \dagger } f , \psi \rangle = \langle f , I^{ ( \chi , j ) } \psi \rangle $, where the former bracket is the inner product of $ L_2(G) $ while the latter is the one of $ \Gamma^{ \chi } $. One can then show that $ I^{ ( \chi , j ) \dagger } $ picks up the $j$-th component of a $\chi$-equivariant function: \begin{equation} ( I^{ ( \chi , j ) \dagger } f ) ( g ) = \sqrt{ d_\chi } f^j ( g ). \label{13.1} \end{equation} \par Next let us turn to the time evolution operator $ U_t $. Observe first that, thanks to the invariance (\ref{7}), the product $ U_t^{ ( \chi , j ) } = I^{ ( \chi , j ) } \, U_t \, I^{ ( \chi , j ) \dagger } $ may be used to define a unitary time evolution projected on $ \Gamma^\chi $. Explicitly, it is given by \begin{equation} ( U_t^{ ( \chi , j ) } f )^i ( g' ) = \int_G dg \int_H dh \sum_{ k=1 }^{ d_\chi } \, \rho_\chi ( h )^{ ik } \, K_G ( g' h , g ; t ) f^k ( g ), \label{13.2} \end{equation} which shows that $ U_t^{ ( \chi , j ) } $ is in fact independent of $ j $, and hence can be written simply as $ U_t^\chi $. From this expression we can deduce the projected propagator $ K_Q^\chi $ acting on $ \Gamma^\chi $ via $( U_t^\chi f ) ( g' )= \int_G dg \, K_Q^\chi ( g' , g ; t ) \, f ( g )$, that is, \begin{equation} K_Q^\chi ( g' , g ; t ) = \int_H dh \, \rho_\chi ( h ) \, K_G ( g' h , g ; t )\ . \label{14} \end{equation} The projected propagator $ K_Q^\chi $ is a map $ K_Q^\chi : G \times G \times \mbox{\boldmath $ R $}^+ \to \mbox{End} ( V_\chi ) $, which is an analogue of (\ref{2}). Note that the summation $ \sum_n \, ( n \in \mbox{\boldmath $ Z $} ) $ with respect to covering points is replaced by the integration $ \int_H dh $ along the fiber as planned, whereas the phase factor $ \omega_n $ is now replaced by the nonabelian weight $ \rho_\chi (h) $. Note also that the composition law $ U_{t+t'}^\chi = U_{t'}^\chi U_t^\chi $ is ensured by the homomorphism $ \rho_\chi (h' h) = \rho_\chi (h') \rho_\chi (h) $ of the representation. The projected propagator $ K_Q^\chi $ has the following properties, \begin{eqnarray} && K_Q^\chi ( g' h , g ; t ) = \rho_\chi ( h )^{ -1 } \, K_Q^\chi ( g' , g ; t )\ , \label{16} \\ && K_Q^\chi ( g' , g h ; t ) = K_Q^\chi ( g' , g ; t ) \, \rho_\chi ( h )\ . \label{17} \end{eqnarray} Thus we see that our path integral on $ Q $ has successfully accommodated the inequivalent quantizations which are labeled by the irreducible representation $ \chi $ of $H$. \section{Inducing the gauge field} Having found the path integral which reproduces the inequivalent quantizations obtained in algebraic approaches, we now move on to examine whether it carries the canonical connection as a gauge field in the form of the (nonabelian) minimal coupling, as we have seen in (\ref{6}) for the case $S^1$. This requires to analyze the local structure of the propagator (\ref{16}) by dividing a path in $ Q $ into small intervals, and for this we need some preparations. \par Recall first that the Haar measure $ dg $ of $ G $ induces the $ G $-invariant measure $ dq = \pi_\ast ( dg ) $ on $ Q $, whereby a function $ \phi : Q \to \mbox{\boldmath $ C $} $ can be integrated as \begin{equation} \int_Q dq \, \phi( q ) = \int_G dg \, \phi( \pi ( g ) ). \label{18} \end{equation} Let $ \{ D_\alpha \} $ be an open covering of $ Q = \cup_\alpha D_\alpha $, $ \{ s_\alpha : D_\alpha \to G \} $ be a set of local sections of the fiber bundle $ \pi : G \to Q $, and $ \{ w_\alpha : Q \to \mbox{\boldmath $ R $} \} $ be a partition of unity associated with the covering $ \{ D_\alpha \} $. Then they give local expressions to various objects: for a $ \chi $-equivariant function $ f $ its pullback is $ f_\alpha = s_\alpha^\ast f = f \circ s_\alpha : D_\alpha \to V_\chi $; the pullback of the projected propagator $ K_Q^\chi $ is a map $ K_{ \alpha \beta }^{ \chi } : D_\alpha \times D_\beta \times \mbox{\boldmath $ R $}^+ \to \mbox{End} ( V_\chi ) $ defined by $ K_{ \alpha \beta }^{ \chi } ( q' , q ; t ) = K_Q^\chi ( s_\alpha( q' ), s_\beta( q ) ; t ) $; and if $ q' \in D_\alpha \cap D_\gamma $ the local expressions are related by $ K_{ \gamma \beta }^\chi ( q' , q ; t ) = \rho_\chi ( t_{ \gamma \alpha } ( q' ) ) K_{ \alpha \beta }^\chi ( q' , q ; t ) $ with a transition function $ t_{ \gamma \alpha } ( q' ) = s_\gamma ( q' )^{-1} s_\alpha ( q' ) $. In terms of these, the time evolution operator reads \begin{equation} ( U_t^\chi f )_\alpha ( q' ) = \sum_\beta \int_Q dq \, K_{ \alpha \beta }^\chi ( q' , q ; t ) \, w_\beta ( q ) f_\beta ( q ). \label{19} \end{equation} Hence the composition law $ U_{ t + t' }^\chi = U_{ t' }^\chi U_{ t }^\chi $ implies \begin{eqnarray} K_Q^\chi ( g'', g ; t + t' ) & = & \int_G dg' \, K_Q^\chi ( g'', g' ; t' ) \, K_Q^\chi ( g' , g ; t ) \nonumber \\ & = & \sum_\alpha \int_{ D_\alpha } dq' \, w_\alpha (q') \, K_Q^\chi ( g'', s_\alpha (q') ; t' ) \, K_Q^\chi ( s_\alpha (q') , q ; t ). \label{20} \end{eqnarray} Inserting intermediate points repeatedly, we obtain \begin{eqnarray} K_Q^\chi ( g_n , g_0 ; t ) & = & \sum_{ \alpha_1 , \cdots , \alpha_{ n-1 } } \int_{ D_{\alpha_{n-1}} } dq_{n-1} \cdots \int_{ D_{\alpha_1} } dq_1 \, w_{ \alpha_{n-1} } ( q_{n-1} ) \cdots w_{ \alpha_1 } ( q_1 ) \nonumber \\ && \qquad \times\, K_Q^\chi ( g_{ n }, s_{ \alpha_{ n-1 } } ( q_{ n-1 } ) ; \epsilon )\, K_Q^\chi ( s_{ \alpha_{ n-1 } } ( q_{ n-1 } ) , s_{ \alpha_{ n-2 } } ( q_{ n-2 } ) ; \epsilon ) \cdots \nonumber \\ && \qquad \times\, K_Q^\chi ( s_{ \alpha_{ 2 } } ( q_{ 2 } ) , s_{ \alpha_{ 1 } } ( q_{ 1 } ) ; \epsilon )\, K_Q^\chi ( s_{ \alpha_{ 1 } } ( q_{ 1 } ) , g_{ 0 } ; \epsilon ), \label{21} \end{eqnarray} where $ \epsilon = t / n $. When two points $ q_k = q( \tau ) $ and $ q_{ k+1 } = q ( \tau + \epsilon ) $ are close enough to be contained in a single patch $ D_\alpha $, one of the factorized propagator becomes \begin{equation} K_{ \alpha \alpha }^\chi ( q ( \tau + \epsilon ) , q ( \tau ) ; \epsilon ) = \int_H dh \, \rho_\chi ( h ( \epsilon ) ) \, K_G ( s_\alpha ( q ( \tau + \epsilon ) ) h ( \epsilon ) , s_\alpha ( q ( \tau ) ) ; \epsilon ), \label{22} \end{equation} where we extend $ h \in H $ to be a smooth function $ h : ( - \epsilon , \epsilon ) \to H $ such that $ h(0) = e $ ($e$ is the identity element of $H$) and $ h( \epsilon ) = h $. Then eq.(\ref{8}) tells that for a short time interval $ \epsilon $, \begin{eqnarray} && K_G ( s_\alpha ( q ( \tau + \epsilon ) ) h ( \epsilon ) , s_\alpha ( q ( \tau ) ) ; \epsilon ) \nonumber \\ && \qquad \quad \approx \exp \left[ i \epsilon \biggl\{ \frac{1}{2} \biggl| \biggl| \Bigl. \frac{d}{ d \epsilon } s_\alpha ( q ( \tau + \epsilon ) ) h ( \epsilon ) \Bigr|_{ \epsilon = 0 } \biggr| \biggr|^2 - V ( q ( \tau ) ) \biggr\} \right], \label{23} \end{eqnarray} with \begin{eqnarray} && \biggl| \biggl| \Bigl. \frac{d}{ d \epsilon } s_\alpha ( q ( \tau + \epsilon ) ) h ( \epsilon ) \Bigr|_{ \epsilon = 0 } \biggr| \biggr|^2 \nonumber \\ & = & \mbox{Tr} \biggl[ \Bigl. h(\epsilon)^{-1} s_\alpha ( q ( \tau ) )^{ -1 } \frac{ d s_\alpha ( q ( \tau ) ) }{ d \tau } h(\epsilon) \Bigr|_{ \epsilon = 0} + \Bigl. h(\epsilon)^{-1} \frac{ d h ( \epsilon ) }{ d \epsilon } \Bigr|_{ \epsilon = 0} \biggr]^2 \nonumber \\ & = & \mbox{Tr} \biggl[ s_\alpha ( q ( \tau ) )^{ -1 } \frac{ d s_\alpha ( q ( \tau ) ) }{ d \tau } + \Bigl. \frac{ d h ( \epsilon ) }{ d \epsilon } h(\epsilon)^{-1} \Bigr|_{ \epsilon = 0 } \biggr]^2 \nonumber \\ & = & \mbox{Tr} \biggl[ P_{ \cal H } \Bigl( s_\alpha ( q ( \tau ) )^{ -1 } \frac{ d s_\alpha ( q ( \tau ) ) }{ d \tau } \Bigr) + \Bigl. \frac{ d h ( \epsilon ) }{ d \epsilon } h(\epsilon)^{-1} \Bigr|_{ \epsilon = 0 } \biggr]^2 \nonumber \\ && + \, \mbox{Tr} \biggl[ P_{ \cal H }^{ \perp } \Bigl( s_\alpha ( q ( \tau ) )^{ -1 } \frac{ d s_\alpha ( q ( \tau ) ) }{ d \tau } \Bigr) \biggr]^2, \label{24} \end{eqnarray} where $ P_{\cal H} $ is a projector from the Lie algebra of $ G $ onto the Lie algebra of $ H $, and $ P_{ \cal H }^{ \perp } $ denotes its orthogonal complement. \par Now, if we take the interval $\epsilon$ small enough, then the contribution from the stationary point of (\ref{24}) with respect to the variation of $ h $ will dominate in the integration $ \int_H dh $ in (\ref{22}). Thus in the limit $\epsilon \rightarrow 0$ the integration may be replaced by the value at the stationary point \begin{equation} \Bigr. \frac{ d h ( \epsilon ) }{ d \epsilon } h(\epsilon)^{-1} \Bigr|_{ \epsilon = 0 } = - P_{ \cal H } \Bigl( s_\alpha ( q ( \tau ) )^{ -1 } \frac{ d s_\alpha ( q ( \tau ) ) }{ d \tau } \Bigr). \label{25} \end{equation} This result may be interpreted that for a small change of the parameter $ q( \tau ) $ in the base manifold $ Q $, the lifted point in the fiber space $ G $ moves along the shortest path, {\it i.e.}, it acquires the smallest change. Now we notice that the right-hand side of (\ref{25}) is nothing but (the pullback of) the canonical connection $ A $, which is just the Maurer-Cartan 1-form $ g^{-1} dg $ projected down to the subalgebra $ {\cal H} $, \begin{equation} A = P_{\cal H} ( g^{-1} dg ). \label{26} \end{equation} It is worth mentioning that this connection is invariant under the $G$-action over the homogeneous space $Q$ and provides various topological solutions of the Yang-Mills equation, for instance, the Dirac monopole and the BPST instanton on $Q = S^2$ and $S^4$, respectively (see, for example \cite{Bais,McMullan}). \par Writing the pullback of $ A $ by the section $ s_\alpha $ as $ A_\alpha = P_{\cal H} ( s_\alpha^{-1} ds_\alpha ) $, we can write the solution of (\ref{25}) as $ h_{\alpha\alpha}[q_{k+1}] = h(\epsilon) = {\cal P}\, \exp [ - \int_{ q_{k} }^{ q_{k+1} } A_{ \alpha } ] $, where the symbol ${\cal P}$ denotes the path-ordering. Using this and gathering scattered pieces, we finally obtain the propagator (\ref{21}) in the desired form, \begin{equation} K_{ \alpha_n \alpha_0 }^\chi ( q_n , q_0 ; t ) = \int_{q_0}^{q_n} [ dq ] \, \rho_\chi ( h_{ \alpha_n \alpha_0 } [ q ] ) \exp \left[ i \int dt \biggl\{ \frac{1}{2} \biggl| \biggl| \frac{ d q }{ d t } \biggr| \biggr|^2 - V ( q ) \biggr\} \right], \label{27} \end{equation} where $ h_{ \alpha_n \alpha_0 } $ is a nonabelian weight factor \begin{eqnarray} && h_{ \alpha_n \alpha_0 } [q] \, = \, t_{ \alpha_n \alpha_{n-1} } ( q_n ) \, {\cal P}\,\exp [ - \int_{ q_{n-1} }^{ q_n } A_{ \alpha_{n-1} } ] \nonumber \\ && \qquad \times \, t_{ \alpha_{n-1} \alpha_{n-2} } ( q_{n-1} ) \, {\cal P}\exp [ - \int_{ q_{n-2} }^{ q_{n-1} } A_{ \alpha_{n-2} } ] \cdots t_{ \alpha_1 \alpha_0 } ( q_1 ) \, {\cal P}\exp [ - \int_{ q_0 }^{ q_1 } A_{ \alpha_0 } ] \label{28} \end{eqnarray} with $ q_k \in D_{k-1} \cap D_{k} $ $ ( k = 1 , \cdots , n ) $ being intermediate points. The factor $ h_{ \alpha_n \alpha_0 }[q] $ is actually a holonomy associated with the path $ q : [ 0, t ] \to Q $, and the above expression (\ref{28}) shows that the gauge field interacts minimally in the nonabelian sense \cite{McMullan}. We therefore reached the path integral (\ref{27}) with (\ref{28}) which precisely reproduces the result found earlier in algebraic approaches~\cite{Landsman,LL,Ohnuki}. \section{Concluding remarks} We considered the path integral on a homogeneous space $ Q = G/H $, and showed that the propagator on $ Q $ can be reduced from the one on $ G $ by integration of redundant degrees of freedom in the fiber direction of $ H $, with a non-trivial weight factor $ \rho_\chi (h) $ multiplied. Being a unitary representation of $ H $, the factor $ \rho_\chi (h) $ preserves the composition law of the propagator, and a different $ \rho_\chi $ leads to a different (inequivalent) quantization. The composition law then allows for a decomposition of the propagator into small intervals, and integration over intermediate points eventually results in the path integral expression. Examination of the propagator at short distance reveals that a gauge field is induced in the path integral in the form of the canonical connection. Thus we have shown that our guiding principle --- `first lift and then project' --- yields the inequivalent quantizations and the induced gauge field correctly. The basic tool used here is essentially the one used in \cite{LL}, where the same path integral expression has been derived from the self-adjoint Hamiltonian through the Trotter formula. In this paper we put an emphasis on the role of the geometry and adopted the guiding principle in order to reach the path integral expression, rather than defining the quantum theory algebraically first. \par Several questions are still left open. One obvious question is how we construct a quantum theory on inhomogeneous spaces. Inhomogeneous spaces often arise in physics, with the one most frequently discussed being a Riemann surface with higher genus. Actually our formulation is not restricted to homogeneous spaces. A more general situation which allows our principle to be employed is the following\footnote{% Such a situation has already been considered by Montgomery~\cite{Montgomery} in investigating geometric properties of induced gauge fields of deformable bodies.}. Let $ P $ be a Riemannian manifold with a metric $ \tilde{g} $ and let a Lie group $ H $ act on $ P $ freely and isometrically. Then the manifold $ M = P/H $ admits an induced metric $ g $, with which the projection $ \pi : P \to M $ defines a principal bundle and a Riemannian submersion. Assume that a propagator in $ P $ is $ H $-invariant, $ K_P ( p'h, ph; t ) = K_P ( p', p; t ) $. Then our formulation of the path integral can be applied straightforwardly. Indeed, the propagator on $M$ can be defined by \begin{equation} K_M^\chi ( p', p; t ) = \int_H dh \, \rho_\chi(h) K_P ( p'h, p; t ), \end{equation} which acts on a $ \chi $-equivariant function $ f : P \to V_\chi $; $ f ( p h ) = \rho_\chi ( h )^{-1} f ( p ) $. When the base space $ ( M, g ) $ is fixed, inequivalent quantizations are classified by choice of the principal bundle $ ( P, M, \pi, H ) $, the lifted metric $ ( P, \tilde{g} ) $ and the representation $ ( H, V_\chi, \rho_\chi ) $. However, this scheme may be too general; we do not have any criterion to choose a specific quantization. In fact, in this scheme the choice of $ ( P, M, \pi, H, \tilde{g} ) $ is equivalent to introduction of an arbitrary gauge field by hand and, as a result, we have no longer a natural explanation of inducing gauge fields. \par In contrast, there exists such a criterion when the base space $ M $ is a homogeneous space $ Q = G/H $. In fact, the invariance under the $ G $-action determines both $ g $ and $ \tilde{g} $ uniquely, and hence the induced gauge field, too. The only remaining arbitrariness is the choice of the representation $ \rho_\chi $, and accordingly there are (infinitely) many inequivalent quantizations. We may therefore conclude that the existence of inequivalent quantizations is the norm when quantizing on a general Riemannian manifold $ ( M, g ) $. If $ M $ admits a transitive action of some isometry group $ G $, then the request of invariance will severely restrict possible quantizations. If $ M $ does not admit such an action, even a self-adjoint momentum operator cannot be defined globally as a generator of the transitive action, and hence in that case we are forced to give up the concept of momentum. \par This last point may be important in realizing the significance of our guiding principle. Indeed, given a homogeneous space $Q = G/H$ there appears no compelling reason, at a glance, to lift it to $G$ and consider quantization there. But this way we can guarantee that there exists a self-adjoint Hamiltonian given by the quadratic Casimir, which in turn ensures unitary time evolution of the system. The existence of such a Hamiltonian is by no means guaranteed for a system whose configuration space is nontrivial. Unfortunately, this is not derived on the sole ground of geometry, and finding such a derivation will be crucial in developing a path integral based on a purely geometric and intuitive principle. \par As a final remark, we add that we have begun a preliminary investigation in two dimensions into the meaning of nontrivial topology in field theories which admit inequivalent quantizations (see, for example~\cite{sigma}). We have however left untouched the higher dimensional cases, not to mention the path integral approach. \section*{Acknowledgments} S.T. wishes to thank T.~Iwai and Y.~Uwano for their encouragement and helpful discussions. I.T. is grateful to D. McMullan for useful advice. This work is supported in part by the Grant-in-Aid for Scientific Research from the Ministry of Education, Science and Culture (No. 07804015). \parskip 0pt
\section{Introduction} Understanding the fermion mass structure has been a goal of particle theorists for some time. In 1978, Froggatt and Nielsen \cite{fn} found that a spontaneously broken family dependent symmetry could naturally explain the large mass ratios among different families of quarks and leptons. Renormalizing experimental data to the Planck scale reveals the order of magnitude estimates to the following ratios \cite{op,rr,br}: \begin{eqnarray} {m_u \over m_t} = {\cal O}(\lambda^8)\;; \qquad\qquad {m_d \over m_b}&=&{\cal O}(\lambda^4)\;; \qquad\qquad {m_e \over m_\tau}={\cal O}(\lambda^4) \nonumber\\ {m_c \over m_t} = {\cal O}(\lambda^4)\;; \qquad\qquad {m_s \over m_b}&=&{\cal O}(\lambda^2)\;; \qquad\qquad {m_\mu \over m_\tau} = {\cal O}(\lambda^2)\;, \label{eqratios} \end{eqnarray} where $\lambda\simeq 0.22$ is the small parameter used in the Wolfenstein's pa\-ra\-me\-tri\-za\-tion \cite{w} of the Cabibbo-Kobayashi-Maskawa (CKM) mixing matrix \cite{c,km}. In the past few years, there has been a revival of theories which predict a mass hierarchy from a spontaneously broken family symmetry. This time the work has been done in the context of supersymmetry (SUSY). The general idea has been used widely in more detailed models with family symmetries that were continuous and discrete, Abelian and non-Abelian, global and local, and with different choices for the symmetry breaking scale \cite{br,ns,ln,ks,ir,dp}. One of the unanswered questions of the original Froggatt-Nielsen model is the origin of the family symmetry breaking. It has been suggested \cite{br,ir} that the supersymmetric versions of the model may be derived from superstring compactification, where spontaneously broken anomalous $U(1)$ gauge symmetries typically occur. In models where the anomalies are canceled by the Green-Schwarz mechanism \cite{gs}, the symmetry breaking scale is slightly below the string scale. Preserving supersymmetry at the high scale determines the vacuum expectation value (VEV) of the symmetry breaking field $\theta$ and the hierarchy parameter $\lambda$, which greatly restricts the theory. The purpose of our work is to find out whether this new constraint can be accommodated in a phenomenologically acceptable model. In this paper, we present models which predict the fermion masses and mixings in a string-inspired framework. Since we do not work with an exact string model, we carry a model-independent analysis as far as possible. In doing so we make the following assumptions: (1) the additional Froggatt-Nielsen symmetry is an anomalous $U(1)$ originating in string theory so that the anomalies are canceled by the Green-Schwarz mechanism; (2) renormalization of couplings and particle masses is done within the framework of the minimal supersymmetric standard model (MSSM); (3) the $U(1)$ symmetry is broken by the VEV of only one field, $\theta$; (4) the Yukawa coupling of $\theta$ to the fermions $f_\theta$ is one. For most of the paper, we assume that the Kac-Moody level, $k_{GUT}$, for the Grand Unified Theory (GUT) group is one, but we keep $k_{GUT}$ as a parameter in most equations. Finally, in the context of a particular string model, the $U(1)$ symmetry breaking field(s) will be known, as will the value of $f_\theta$ and $k_{GUT}$. The outline of the paper is as follows. In section~\ref{secreview}, we summarize the main features of the Froggatt-Nielsen models. In section~\ref{secsusy} we discuss the implications of unbroken supersymmetry for the value of the mass hierarchy parameter, $\lambda$. This is followed by section~\ref{secgsmech}, which gives the background and some important facts about the Green-Schwarz mechanism. Section~\ref{secanomaly} is dedicated to the anomalies of the model. We show there how $\lambda$ depends only on the $X$ charges of the standard model fields. In section~\ref{secpq} we present some constraints on those charges following from relations (\ref{eqratios}). We illustrate these constraints in section~\ref{secexamples} by working out in detail a few simple models. Section~\ref{secconclusions} summarizes our results. \section{Froggatt-Nielsen Models} \label{secreview} Originally, Froggatt and Nielsen proposed \cite{fn} that a flavor-dependent symmetry be broken by the VEV of an additional scalar field, $\theta$, which would be a singlet of the standard model gauge groups. Their idea also assumed a set of heavy ``mirror quarks'', analogous to the standard model quarks, with a spectrum of charges under the horizontal symmetry. Mass matrices would then arise through effective Yukawa interactions resulting from Feynman diagrams such as that in figure~\ref{figfeynman1}.% \begin{figure} \begin{center} \newsavebox\dashed \begin{picture}(300,100) \put(5,0){\line(1,0){290}} \sbox{\dashed}{\multiput(0,0)(0,10){8}{\line(0,1){5}}% \put(-4.58,72){$\times$}\put(-7.5,85){$\langle\theta\rangle$}} \multiput(50,0)(0,10){8}{\line(0,1){5}}\put(50,80){\line(0,1){3}} \put(45,85){$H$} \multiput(90,0)(40,0){5}{\usebox\dashed} \multiput(30,0)(40,0){7}{\vector(1,0){0}} \put(20,10){$+2$} \put(60,10){$+2$} \put(100,10){$+1$} \put(147,10){0} \put(180,10){$-1$} \put(220,10){$-2$} \put(260,10){$-3$} \end{picture} \end{center} \caption{Tree diagram leading to effective Yukawa interactions. Above the solid line are $U(1)$ charges of quarks and mirror quarks.} \label{figfeynman1} \end{figure} In figure~\ref{figfeynman1}, we show an example, where the $X$ charge assignments of the quarks are written above each quark line. On one side of the diagram, we have a left-handed quark doublet with charge $+2$. If we assign a charge of $-3$ to the right-handed quark, then there must be five $\theta$ interactions, with five mirror quarks of charges $+2$, $+1$, 0, $-1$, $-2$ in between. The first mirror quark of charge $+2$ interacts with the standard model Higgs doublet and the quark doublet to conserve $SU(2)$ symmetry. Assuming a common mass $M$ for the mirror quarks, and a common Yukawa coupling $f_\theta$ of the $\theta$ field to all the quarks, $\langle\theta\rangle$ should take a value such that $\lambda \sim f_\theta \langle\theta\rangle/M$. The mass term resulting from fig.~\ref{figfeynman1} would be \begin{equation} \label{eqquarka} f_u \bar u_j Q_i H \left(f_\theta {\langle\theta\rangle\over M} \right)^5. \end{equation} In a realistic model, the charges must be assigned so that all the mass matrix eigenvalues agree with the relations~(\ref{eqratios}) above. Here, we do not make the Froggatt-Nielsen assumptions, and thus do not require mirror quarks. We assume that the flavor symmetry is a gauged $U(1)$ symmetry, labeled by $X$, left over from string compactification. We expect the action to contain all terms consistent with charge conservation. Such terms appear due to string tree diagrams; therefore, the effective Yukawa coupling $f_\theta$ will be a product of the string coupling constant $g_s$ and other terms of order unity. Not knowing the details of the model, we cannot evaluate $f_\theta$, and here assume $f_\theta=1$. In order to demonstrate the generation of mass terms, we give an explicit example. We do not make any further assumptions about the physical mechanism. First, we define the $X$ charges to be $q_{Qi}$, $q_{ui}$, $q_{di}$, $q_{Li}$, and $q_{ei}$ for the left-handed quark doublets, the left-handed up-type antiquarks, the left-handed down-type antiquarks, the left-handed lepton doublets and the left-handed positrons ($i$\/ is the family index). Also, we define $q_H$ to be the sum of the $X$ charges for the two Higgs doublets of the supersymmetric standard model. We then consider all bilinear fermion terms that conserve $X$ charge. A typical up-type quark term would be \begin{equation} \label{eqquark} \left({\bf Y}_u\right)_{ij}= f_u \bar u_j Q_i H_1 \left({\langle\theta\rangle\over M}\right)^{q_{Qi}+q_{uj}+q_{H1}} \end{equation} The entire Yukawa mass matrix then follows: \begin{equation} \label{eqmatrix} {\bf Y}_u= f_u \,\lambda^{q_{H1}} \left( \begin{array}{ccc} \lambda^{q_{Q1}+q_{u1}} &\lambda^{q_{Q1}+q_{u2}} &\lambda^{q_{Q1}+q_{u3}}\\ \lambda^{q_{Q2}+q_{u1}} &\lambda^{q_{Q2}+q_{u2}} &\lambda^{q_{Q2}+q_{u3}}\\ \lambda^{q_{Q3}+q_{u1}} &\lambda^{q_{Q3}+q_{u2}} &\lambda^{q_{Q3}+q_{u3}} \end{array} \right). \end{equation} \section{Implications of Unbroken Supersymmetry} \label{secsusy} The basic premise of this work is the assumption of a deeper connection between string theory and supersymmetric models with spontaneously broken family symmetries. By assuming such a connection, we can ``borrow'' a $U(1)$ gauge symmetry left over from string compactification. We begin with $N=1$ global supersymmetry and the scalar potential: \begin{equation} V={1\over2}\sum_\alpha (D^\alpha)^2 + \sum_i |F_i|^2. \end{equation} Here $F_i = \partial W/\partial\phi_i$. We do not specify the superpotential $W$; thus, we will not be able to predict the VEV's of all of the fields in the model. The gauge $D$ term is given by $D^\alpha=g_{(\alpha)}\sum_{ij}\phi_i^\ast(T^\alpha)_{ij}\phi_j$, where $\phi_i$ are the matter chiral superfields, $T^\alpha$ are the generators of the gauge group, and $g_{(\alpha)}$ are the gauge couplings. For an anomalous $U(1)$ gauge group, the corresponding $D$ term will be modified by a Fayet-Iliopoulos term. Its magnitude has been calculated in string theory \cite{ad,di,ds} on the assumption that the anomalies are canceled by the Green-Schwarz mechanism \cite{gs}: \begin{equation} \label{eqDterm} D={g_s M_s^2\over192\pi^2} {\,\rm tr\,} Q + \sum_i q_i |\phi_i|^2, \end{equation} where $g_s$ is the renormalized string coupling constant, $M_s$ is the string scale, and ${\,\rm tr\,} Q \equiv \sum_i q_i$ is the sum of the $U(1)$ charges of all the particles. (See also \cite{fi} for a clear exposition.) In the model we are studying, the $U(1)_X$ family symmetry gauge group is anomalous, has origins in string theory and requires a $D$ term given by eq.~(\ref{eqDterm}). The supersymmetric vacuum requires $\langle F_i\rangle=0$ and $\langle D^\alpha\rangle=0$. If the $X$ charges of all the particles are of the same sign, then, according to eq.~(\ref{eqDterm}), it is impossible to preserve supersymmetry. We therefore require that charges with both signs be present. (Fortunately, this is typically the case in string compactifications.) As a convention, we give the $\theta$ field a negative charge. In section~\ref{secexamples}, we will consider both the case in which all the standard model matter fields have positive $X$ charge and the case in which they can have charges of either sign, subject to the condition ${\,\rm tr\,} Q>0$. We have to assume that none of the fields charged under the standard model (SM) gauge groups can develop vacuum expectation values---otherwise color or electroweak symmetry would be broken at the high scale $\sim M_s$. The problem of flat directions here does not differ from the problem of flat directions in the MSSM. We require $\langle D^\alpha\rangle=0$ for each gauge factor separately; for each of the SM gauge groups this condition involves only SM fields, so that the flat directions will be the same as in the MSSM. For $U(1)_X$, setting $\langle D\rangle=0$ merely determines $\langle\theta\rangle$ and does not constrain MSSM fields. Hence, just as in the MSSM, we have to rely on the superpotential to lift the flat directions. The most important implication of eq.~(\ref{eqDterm}) for this work is that preserving supersymmetry determines the VEV of the $\theta$ field \begin{equation} \label{eqsqrt} {\langle\theta\rangle\over M_s} = \sqrt{{g_s\over192\pi^2} {{\,\rm tr\,} Q \over |q_\theta|}}\;. \end{equation} If we assign $X$ charges to all the fields and use $\theta$ as the family symmetry breaking field in the Froggatt-Nielsen scheme, we obtain a prediction for the Yukawa mass matrices. The question we attempt to answer in this paper is: can we find a set of charges for all the standard model fields, consistent with the requirements of anomaly cancellation, that will predict phenomenologically viable powers of $\lambda$ (much like the work by Ib\'a\~nez and Ross \cite{ir}, but without the assumption of left-right symmetry) \ {\em and\/} \ predict a phenomenologically viable value of $\lambda\simeq0.22\,$? We shall see that this is possible as long as $f_\theta$ is not very different from unity. Our results, as can be seen from eq.~(\ref{eqsqrt}), are rather sensitive to the value of the string coupling constant, $g_s$. At the unification scale, we use the tree-level relation \cite{k,g} \begin{equation} \label{eqkgut} 1/g_s^2=k_{GUT}/g_{GUT}^2, \end{equation} where $k_{GUT}$ is the Kac-Moody level for corresponding GUT gauge group algebra. Here, we take $k_{GUT}=1$. We then use a typical value $\alpha_{GUT} = {g_{GUT}}^2/4\pi \simeq 1/25$ to get $g_s \simeq 0.7$. Substituting into eq.~(\ref{eqsqrt}) with $q_\theta=-1$, we obtain \begin{equation} \lambda = \langle\theta\rangle/M_s = 1.92 \times 10^{-2} \, \sqrt{{\,\rm tr\,} Q}. \end{equation} \section{Green-Schwarz Anomaly Cancellation} \label{secgsmech} In chiral theories, it is necessary to consider the problem of quantum anomalies. These anomalies to classical symmetries are dangerous in that they prevent the existence of gauge theories. In this section we discuss a method of removing anomalies that may arise with a new $U(1)_X$ gauge symmetry. We start with $U(1)$ chiral transformations on all fermions: \begin{eqnarray} \label{eqtransf} \Psi(x)&\to& \exp\left[-iq\gamma_5 \Theta(x)\right]\Psi(x)\\ \bar\Psi(x)&\to& \bar\Psi(x) \exp\left[-iq\gamma_5 \Theta(x)\right], \nonumber \end{eqnarray} where $q$ is the charge of each fermion. Since the path integral measure is not invariant under the transformation, we obtain new terms beyond the usual current divergence \cite{f}. The difference can be expressed as a change in the Lagrangian: \begin{equation} \label{eqLag} {\cal L}\to {\cal L} - \Theta(x)\!\!\!\!\sum_{i=1,2,3,X}\!\!\!\! C_i F_i \tilde F_i - \Theta(x) \left(\partial_\mu j_5^\mu\right) \end{equation} where the sum is over the standard model gauge groups and $U(1)_X$. The coefficient $C_1 = {\,\rm tr\,}\left[Q (Y/2)^2\right]$ is the mixed $U(1)_X \left(U(1)_Y\right)^2$ anomaly, $C_X = {\,\rm tr\,} Q^3$ is the $\left(U(1)_X\right)^3$ anomaly, and $C_{2,3} = {1\over2} {\rm tr}_{2,3}\,Q$ are the $U(1)_X \left(SU(2)_L\right)^2$ and $U(1)_X \left(SU(3)_c\right)^2$ anomalies. (The trace $\,{\rm tr}_{2,3}\,Q$ is over fermions with $SU(2)_L$ and $SU(3)_c$ charge, respectively.) In some cases we can choose the charges so that all the anomaly coefficients are zero, but here we examine the possibility of canceling the anomalous term with another of opposite sign \cite{gs}. There is another mixed anomaly with $U(1)_Y$, the $(U(1)_X)^2\, U(1)_Y$, but it does not fit into the discussion. It results from a different transformation than that in eq.~(\ref{eqtransf}), where $q$ would be the $U(1)_Y$ instead of the $U(1)_X$ charge. In the Green-Schwarz mechanism, anomalies are canceled through an additional field. In string theory, the antisymmetric tensor $B_{\mu\nu}$ naturally serves this purpose. In four dimensions \cite{hn}, we can replace $H=dB$ with its dual, which is the derivative of the axion field $\Phi$: \begin{equation} \label{eqdual} dB = *d\Phi. \end{equation} This field couples to the gauge groups in the following way: \begin{equation} \label{eqetacoupl} \Phi\!\!\!\!\sum_{i=1,2,3,X}\!\!\!\! k_i F_i \tilde F_i, \end{equation} where $k_i$ are the Kac-Moody levels of the corresponding gauge algebra. For the $U(1)_X$ gauge transformation, \begin{equation} \label{eqXgauge} A_X^\mu \rightarrow A_X^\mu + \partial^\mu\Theta(x) \end{equation} $\Phi$ follows the transformation \begin{equation} \label{eqetagauge} \Phi\rightarrow\Phi + \Theta(x) \,\delta_{GS}. \end{equation} Therefore, we can remove quantum anomalies through a gauge transformation if $\delta_{GS}=C_1/k_1=C_2/k_2=C_3/k_3=C_X/k_X$. For a more detailed discussion, we refer the reader to the paper by Ib\'a\~nez \cite{i}. So far, we have ignored gravity, but the conclusions do not change. We must only cancel one additional anomaly, $C_{\rm grav} R \tilde R$, through a gauge transformation on one additional coupling $k_{\rm grav} \Phi R \tilde R$. Finally, we have: \begin{equation} \label{eqckratio} C_1/k_1=C_2/k_2=C_3/k_3=C_X/k_X=C_{\rm grav}/k_{\rm grav}. \end{equation} In this paper, we take $k_1=5/3$ and $k_2=k_3=k_{\rm grav}=1$, and we do not use $k_X$, so that \begin{equation} \label{eqcratio} C_1:C_2:C_3:C_{\rm grav} = {5\over3}:1:1:1. \end{equation} \section{Quantum Anomalies} \label{secanomaly} We can now apply the results of the previous section to the case at hand. The mixed anomalies with the standard model gauge groups are \begin{eqnarray} \label{eqC123} C_1 &=& {1\over6} (q_Q + 8q_u + 2q_d +3q_L +6q_e +3q_H) \nonumber\\ C_2 &=& {1\over2} (3q_Q + q_L + q_H) \\ C_3 &=& {1\over2} (2q_Q + q_u + q_d), \nonumber \end{eqnarray} where $q_Q = \sum_{i=1}^3 q_{Qi}$, etc., but $q_H = q_{H1}+q_{H2}$ is the sum of the $U(1)_X$ charges of the two Higgs doublets. Our calculations do not include the right-handed neutrinos, but since we allow for the existence of additional particles with $X$ charge that are singlets under the standard model, our results do not depend on the existence and $X$ charge assignments of $\nu_R$. The gravitational anomaly is given by \begin{equation} \label{eqcgrav} C_{\rm grav} = {1\over24} \sum_{\mbox{\scriptsize all particles}}\!\!\!\!\!\! q_i = {1\over24} (6q_Q + 3q_u + 3q_d + 2q_L + q_e + 2q_H + q_\theta + q_X), \end{equation} where $q_X$ is the sum of the $U(1)_X$ charges of any additional fields which are singlets under the standard model. We are not excluding such fields, and we cannot evaluate $C_{\rm grav}$ directly. However, because we are using the Green-Schwarz mechanism, we know that $C_{\rm grav}$ must be in the correct proportion~(\ref{eqcratio}) to the other anomalies. {}From the expressions (\ref{eqC123}) and (\ref{eqcgrav}) we then obtain \begin{equation} \label{eq123g} q_\theta + q_X = 18q_Q + 8q_u + 7q_d = 14C_3 + 4q_Q + q_u. \end{equation} With the assumption that $\theta$ is the only field with a negative $U(1)_X$ charge, we see immediately that we {\em require\/} the extra fields with no standard model interactions to balance eq.~(\ref{eq123g}) with a large positive contribution $q_X$. If we allow the quarks to have negative charges, this is no longer true. Even then, in section~\ref{secpq} we shall see that $C_3 \sim 9$ in phenomenologically interesting models, so that the typical model will require $q_X>0$. The additional fields responsible for $q_X\not=0$ prevent us from calculating the cubic anomaly $\left(U(1)_X\right)^3$, which then does not impose any constraints on the model. We simply assume that the charges of the extra fields are such that it is canceled: \begin{equation} C_X = \!\!\!\! \sum_{\mbox{\scriptsize all particles}} \!\!\!\!\!\! q_i^3. \end{equation} On the other hand, the mixed anomaly $\left(U(1)_X\right)^2 \, U(1)_Y$, \begin{equation} C_{YXX} = \!\!\!\! \sum_{\mbox{\scriptsize all particles}} \!\!\!\!\!\! Y_i q_i^2 = \sum_{i=1}^3 \left( q_{Qi}^2 - 2q_{ui}^2 + q_{di}^2 - q_{Li}^2 + q_{ei}^2 \right) - q_{H1}^2 + q_{H2}^2, \end{equation} depends only on the charges of the standard model particles and cannot be canceled by the Green-Schwarz mechanism. For each charge assignment we have to check that $C_{YXX}=0$. The equality of $C_3$ and $C_{\rm grav}$, eq.~(\ref{eqcratio}), is crucial in this paper. Section~\ref{secsusy}, eq.~(\ref{eqsqrt}) gave us a prediction for the hierarchy parameter $\lambda$ in terms of ${\,\rm tr\,} Q$, the sum of the $U(1)_X$ charges of all the particles in the model. That sum, according to eq.~(\ref{eqcgrav}), is proportional to the gravitational anomaly. Because $C_{\rm grav}=C_3$, every charge assignment for the standard model fields (in fact, for the quark fields alone) results in a prediction \begin{equation} \label{eqc3l} \lambda = \sqrt{g_s/8\pi^2}\,\sqrt{C_3} \simeq 0.094\sqrt{C_3}. \end{equation} \section{Determinants of the Mass Matrices} \label{secpq} The product of the determinants of up- and down-quark mass matrices will give us an important constraint. From eq.~(\ref{eqratios}), it is immediately seen that \begin{equation} \label{eqlam} \prod_{\mbox{\scriptsize all quarks}} \!\!\!\!\! m_q \sim f_u^3 f_d^3\,\lambda^{18}. \end{equation} This should be compared to the product of the determinants of the Yukawa matrices predicted by the model. Writing ${\bf Y}_u$ in the form (\ref{eqmatrix}), we see that every term in the determinant is of the order $f_u^3 \,\lambda^{q_Q+q_u+3q_{H1}}$. Similarly, every term in $\det {\bf Y}_d$ is of the order $f_d^3 \,\lambda^{q_Q+q_d+3q_{H2}}$. The two taken together give \cite{n} \begin{equation} \label{eqdet1} \prod_{\mbox{\scriptsize all quarks}} \!\!\!\!\! m_q = \left| \det{\bf Y}_u \right| \: \left| \det{\bf Y}_d \right| \sim f_u^3 f_d^3 \,\lambda^{2q_Q + q_u + q_d + 3q_H}. \end{equation} Now, $q_H=0$ if a $\mu$ term $\mu H_1 H_2$ is to be allowed and not suppressed. (For alternatives, see \cite{n,js}. We note that a small change in $q_H$ can be easily accomodated, as it will not change the predicted mass ratios or mixings.) Then, using eq.~(\ref{eqC123}), we are left with \begin{equation} \label{eqdet2} \prod_{\mbox{\scriptsize all quarks}} \!\!\!\!\! m_q \sim f_u^3 f_d^3 \,\lambda^{2 C_3}. \end{equation} {}From equations (\ref{eqlam}) and (\ref{eqdet2}), we see that \begin{equation} \label{eqc318} \lambda^{18} \sim \lambda^{2C_3}, \end{equation} or \begin{equation} \label{eqc39} C_3\simeq9, \end{equation} as was found earlier by Bin\'etruy and Ramond \cite{br}. This is true whether or not there are texture zeros, provided that neither determinant is zero. For $C_3=9$, eq.~(\ref{eqsqrt}) gives $\lambda=0.28$. The above reasoning assumes that $\lambda$ is fixed at about $0.22$---the mass ratios (\ref{eqratios}) come from experiment, not from assumptions about hierarchy. In this paper, we derive the hierarchy parameter $\lambda$ from supersymmetry. Taking $\lambda$ as predicted by eq.~(\ref{eqc3l}), we have to replace (\ref{eqc318}) by \begin{equation} \label{eqc3f} 0.22^{18} \sim \left(0.094\sqrt{C_3}\right)^{2C_3}. \end{equation} Solving this, we get $C_3 \simeq 12.5$ and $\lambda\simeq0.33$, rather than $0.22$. This value of $\lambda$ will restrict the number of solutions because first order calculations predict a dependency of the Cabibbo angle on $\lambda$. If Yukawa matrices are given in terms of powers of $\lambda$, so will, to leading order, the CKM matrix \cite{fn}. The experimental uncertainty on the average value of the Cabibbo angle \cite{lr,gh} \begin{equation} |{\bf V}_{12}|=0.2205\pm0.0018 \end{equation} is very small. Keeping in mind that, as noted by Olechowski and Pokorski \cite{op}, $|{\bf V}_{12}|$ is almost invariant (it changes by less than 0.1\%) when renormalized from $M_W$ to $M_{GUT}$, we will always try to keep close to $\lambda\simeq0.22$. In order to remedy the solution to eq.~(\ref{eqc3f}), we will examine the assumptions that play a significant role since the equation itself is robust. It is robust because $C_3$ is related to the exponent of a small parameter, so a small change in $C_3$ would change the determinants by orders of magnitude. With $\lambda\simeq0.22$, we estimate that unless the order unity factors in the Yukawa matrices all conspire to shift the balance in one direction, they could increase $C_3$ by as much as two or three. One should also note that there are no top or bottom Yukawa couplings in eq.~(\ref{eqc318}), so the result is independent of $\tan \beta$. One assumption that affects the value of $C_3$ more significantly is $f_\theta=1$. If, for example, $f_\theta$ were $0.78$, then the expected and calculated values of $\lambda$ would be reconciled. Another assumption that can be relaxed concerns the values of the Kac-Moody levels. If, instead, $k_2=k_3=k_{GUT}=2$ while $k_{\rm grav}=1$, then both eq.~(\ref{eqkgut}) relating $g_s$ and $g_{GUT}$ and eq.~(\ref{eqckratio}) relating $C_3$ and $C_{\rm grav}$ must change. (Models with $k_{GUT}=2$ have been increasingly popular with string theorists \cite{l,cc,af}.) The final relation for $\lambda$, eq.~(\ref{eqc3l}), becomes \begin{equation} \label{eqc3lkap} \lambda = f_\theta \sqrt{g_{GUT}\, C_3\over8\pi^2\sqrt{k_{GUT}}} \simeq 0.079 f_\theta \sqrt{C_3}. \end{equation} Now, $\lambda\sim0.24$ when $C_3 = 9$. Lastly, we note that $\lambda$ also depends on the value of $g_s$ by the above equation (for both $k_{GUT}=1$ and $k_{GUT}=2$). However, we know that $f_\theta$ has a linear dependence on $g_s$, so that \begin{equation} \label{eqfg} \lambda\propto g_s^{3/2}. \end{equation} In order to attain a value of $\lambda\sim0.22$, $g_s$ would have to be reduced from $0.7$ to $0.59$. This would require $\alpha_{GUT}=1/36$, which is too low according to most models. \section{Detailed Examples} \label{secexamples} We are now ready to examine in detail the $U(1)_X$ charge assignments, which, subject to the constraints discussed in sections~\ref{secanomaly} and \ref{secpq}, let us calculate fermion mass matrices. We would then compare the quark and lepton masses with relations~(\ref{eqratios}) and demand a phenomenologically viable CKM matrix. Ideally, among all the possible charge assignments we would find at least one that satisfies all the constraints and predicts masses and mixings within experimental bounds. Although the fifteen charges of the quark and lepton fields may seem like many free parameters, they are in fact overconstrained. If we demand $\lambda=0.22$, equations (\ref{eqc3l}) and (\ref{eqdet2}) become two independent predictions for~$C_3$. {\it A priori\/} it is not obvious that the two numbers should even be of the same order of magnitude. When $f_\theta$ and $k$ are taken into account, in the context of a particular string compactification, the two predictions will be more than just order of magnitude estimates. If we do {\em not\/} require $\lambda=0.22$, then we have to be able to produce the correct Cabibbo angle from a texture given in terms of powers of the calculated $\lambda$. Furthermore, the charges are integers, and they are constrained by the mixed anomalies. One of the constraints is non-linear. It is not guaranteed that there will be any solution, much less that it will correspond to realistic masses and mixings. \subsection{Positive Charges for Matter Fields} \label{secpositive} We begin the search for a detailed model by assuming that all the standard model fields have nonnegative $U(1)_X$ charges. We also assume, following Bin\'etruy and Ramond \cite{br}, that because of a possible $\mu$ term, the Higgs doublets have zero $X$ charge. We, therefore, form the Yukawa matrices \begin{equation} \label{eqYuk} {\bf Y}_u = f_u \left( \begin{array}{ccc} \lambda^{(q_{Q1}+q_{u1})/|q_\theta|} & \lambda^{(q_{Q1}+q_{u2})/|q_\theta|} & \lambda^{(q_{Q1}+q_{u3})/|q_\theta|} \\ \lambda^{(q_{Q2}+q_{u1})/|q_\theta|} & \lambda^{(q_{Q2}+q_{u2})/|q_\theta|} & \lambda^{(q_{Q2}+q_{u3})/|q_\theta|} \\ \lambda^{(q_{Q3}+q_{u1})/|q_\theta|} & \lambda^{(q_{Q3}+q_{u2})/|q_\theta|} & \lambda^{(q_{Q3}+q_{u3})/|q_\theta|} \end{array} \right) \end{equation} for the up sector, and similarly for the down and lepton sectors. Eq.~(\ref{eqYuk}) is an order of magnitude relationship, so that each of the entries will be multiplied by a number of order unity. The factors of order unity will be necessary to introduce CP violation, which we are ignoring in this paper. We would like to make $|q_\theta|$ the smallest charge unit, i.e.\ $q_\theta=-1$ (in our convention all $X$ charges are integers). That, however, does not lead to phenomenologically acceptable mass matrices: every row of matrix (\ref{eqYuk}) is equal to $\lambda^{\mbox{\scriptsize some power}}\times (\mbox{some other row})$. Similarly, there is only one independent column. The mass matrix, with two zero eigenvalues, does not reproduce the observed mass hierarchy even qualitatively. Although the factors of order unity multiplying the entries in ${\bf Y}_u$ will in general move $\;\det {\bf Y}_u\;$ away from zero, fermion masses and mixings would then very strongly depend on those unknown factors and not on the properties of the model we are trying to investigate. Such a model may be realized in nature, however, we do not know those factors. We will therefore set them to one in this and the following section, and impose conditions to avoid zero determinants. We will find that those conditions are too rigid to produce a realistic example. Since they are not based on physical principles, in section~\ref{sectf} we make an arbitrary but limited choice of the ``texture factors'' to see how much can be achieved by eliminating the artificial constraints. Without texture factors, any matrix of the form ${\bf Y}_{ij}=\lambda^{a_i+b_j}$ will have rank one. A possible solution is to use matrices with texture zeros: $({\bf Y}_u)_{ij}=\lambda^{(q_{Qi}+q_{uj})/|q_\theta|}$ when $(q_{Qi}+q_{uj})/|q_\theta|$ is a nonnegative integer, zero otherwise. If the sum of the charges $q_{Qi}+q_{uj}$ is not a multiple of $|q_\theta|$, the corresponding term in the Lagrangian is forbidden by charge conservation. If it is negative, it could only be matched by a power of $\theta^\ast$. This is impossible in supersymmetric theories because the superpotential must be holomorphic in $\theta$. With positive charges for the standard model fields, $q_\theta=-1$ does not allow texture zeros. We find that $q_\theta=-2$ is also not enough. The pattern of the texture zeros depends only on the remainder from the division of $q_{Qi}+q_{uj}$ by $|q_\theta|$. Of the three numbers $q_{u1}$, $q_{u2}$, $q_{u3}$, either at least two will be even or at least two will be odd. At least two columns of ${\bf Y}_u$, having the same pattern of texture zeros, will be proportional. The result is a matrix of rank at most two. In order to obtain a non-singular matrix we must have all $q_{Qi}$ different $(\mbox{mod }|q_\theta|)$ and all $q_{ui}$ different $(\mbox{mod }|q_\theta|)$. That requires $|q_\theta|\ge3$. In addition, one can see that no row and no column can have more than one nonzero element. The only non-singular mass matrices we can get in this model are rather sparse: they have six texture zeros. Even without the assumption of left-right symmetry \cite{rr} there are few possibilities: \begin{eqnarray} \label{eqmat} \left(\begin{array}{ccc} a & 0 & 0 \\ 0 & b & 0 \\ 0 & 0 & c \end{array}\right), \qquad & \left(\begin{array}{ccc} 0 & c & 0 \\ b & 0 & 0 \\ 0 & 0 & a \end{array}\right), \qquad & \left(\begin{array}{ccc} a & 0 & 0 \\ 0 & 0 & b \\ 0 & c & 0 \end{array}\right), \\ \left(\begin{array}{ccc} 0 & 0 & b \\ 0 & a & 0 \\ c & 0 & 0 \end{array}\right), \qquad & \left(\begin{array}{ccc} 0 & a & 0 \\ 0 & 0 & b \\ c & 0 & 0 \end{array}\right), \qquad & \left(\begin{array}{ccc} 0 & 0 & c \\ a & 0 & 0 \\ 0 & b & 0 \end{array}\right). \nonumber \end{eqnarray} In addition, all the above textures can be reduced to a diagonal form by a permutation of columns only. Permutations of columns of a Yukawa mass matrix can be written as a multiplication on the right by a unitary permutation matrix, and it is easy to show that they do not change the resulting masses or mixings. (Permutations of rows, which correspond to multiplication on the left, will not change masses or mixings as long as they are done simultaneously in the up and down sector.) We conclude that the masses and mixings will be the same as if all mass matrices were diagonal, i.e.\ there will be no flavor mixing. Can we make things better by giving up some of the assumptions? In the following section we will let all particles have charges of either sign. Here we only note that if we allow the Higgs doublet to have a positive charge (at the expense of the $\mu$ term), we are merely shifting the charges of the up, down or lepton sector without changing any of the conclusions. If the Higgs charge is a multiple of $|q_\theta|$, the textures do not change. Otherwise, the texture zeros will change their positions, but we will still have exactly one nonzero entry in each row and each column. It should be stressed again that our failure to find a workable example of a model with only positive charges for the standard model fields does not in any way rule out such models. Our conclusion here is that we cannot obtain an acceptable mass matrix to study unless we know the exact factors of order unity in front of the powers of $\lambda$ in (\ref{eqYuk}). It would be interesting to come back to this exercise when we can make an informed choice of those factors. \subsection{Allowing Negative Charges} \label{secnegative} We now turn to the analysis of the model in which the matter fields are allowed to have negative as well as positive charges. That will be a source of texture zeros, and give us more flexibility in constructing the mass matrices. We need to find the conditions necessary to obtain a non-singular matrix, and start by reordering the quarks and leptons so that $q_{Q1} \ge q_{Q2} \ge q_{Q3}$, $q_{u1} \ge q_{u2} \ge q_{u3}$, etc. This is the same as permuting the rows and columns of the mass matrix; it can only change the determinant of {\bf Y} by a sign, and will not change the masses or mixings. Now in eq.~(\ref{eqYuk}) we put a texture zero wherever $q_{Qi}+q_{uj}<0$. Keeping in mind that no two columns and no two rows can have the same pattern of texture zeros, we are left with \begin{equation} \label{eqYukneg} {\bf Y}_u = f_u \left( \begin{array}{ccc} \lambda^{q_{Q1}+q_{u1}} & \lambda^{q_{Q1}+q_{u2}} & \lambda^{q_{Q1}+q_{u3}} \\ \lambda^{q_{Q2}+q_{u1}} & \lambda^{q_{Q2}+q_{u2}} & 0 \\ \lambda^{q_{Q3}+q_{u1}} & 0 & 0 \end{array} \right). \end{equation} We are assuming $q_\theta=-1$ in this section. The texture (\ref{eqYukneg}) is rich enough to ask whether it is possible to obtain realistic fermion masses and mixings. To answer this question, we have done a computerized search by trying out all possible charge assignments for the quarks and leptons (in a range from $-10$ to 10), imposing the anomaly constraints, calculating the fermion masses and aiming to be as close to the ratios (\ref{eqratios}) as possible. (We were limited in how close we could get by the relations (\ref{eqc3l}) and (\ref{eqdet2}) between the determinants of the quark mass matrices and the hierarchy parameter $\lambda$.) For those charge sets that produced the best fermion masses, we then computed the CKM matrix. While none of the results reproduce experimental data, we have found many that were not unreasonable. Below we give an example. For $g_s=0.7$ and the following $X$ charges, $$ \begin{tabular}{c|rrrrr} $i$& $q_{Qi}$ & $q_{ui}$ & $q_{di}$ & $q_{Li}$ & $q_{ei}$ \\ \hline 1 & 9 & 9 & 7 & 1 & 10 \\ 2 & 7 & 2 & $-4$ & $-7$ & 9 \\ 3 & $-4$ & $-9$ & $-9$ & $-10$ & 3 \end{tabular} $$ we have $C_3=10$ and $\lambda=0.30$, which results in the fermion mass ratios \begin{eqnarray*} \rule[-2pt]{0pt}{14pt} {m_u \over m_t} = 1.8\times10^{-5} \;, \qquad\quad {m_d \over m_b} &=& 2.6\times10^{-2} \;, \qquad\quad {m_e \over m_\tau} = 7.9\times10^{-3} \;, \\ \rule{0pt}{18pt} {m_c \over m_t} = 2.3\times10^{-2} \;, \qquad\quad {m_s \over m_b} &=& 2.6\times10^{-2} \;, \qquad\quad {m_\mu \over m_\tau} = 8.9\times10^{-2} \;, \\ \rule[2pt]{0pt}{16pt} {m_t \over f_u} = 1.0 \;, \qquad\qquad\qquad\; {m_b \over f_d} &=& 1.0 \;, \qquad\qquad\qquad\; {m_\tau \over f_d} = 1.0 \;, \end{eqnarray*} and the CKM matrix \begin{equation} \label{eqV} {\bf V} = \left( \begin{array}{ccc} 0.96 & 0.27 & 6.2\times10^{-5} \\ -0.27 & 0.96 & 2.0\times10^{-10} \\ -6.0\times10^{-5} & -1.6\times10^{-5} & 1.0 \end{array} \right). \end{equation} The 90\% confidence experimental limits on the magnitude of the CKM matrix elements \cite{rev}, renormalized to the GUT scale \cite{op}, are \begin{equation} \left( \begin{array}{ccc} 0.9747\mbox{ to } 0.9759 & 0.218 \mbox{ to } 0.224 & 0.001 \mbox{ to } 0.003 \\ 0.218 \mbox{ to } 0.224 & 0.9738\mbox{ to } 0.9752 & 0.021 \mbox{ to } 0.032 \\ 0.002 \mbox{ to } 0.010 & 0.020 \mbox{ to } 0.032 & 0.9995\mbox{ to } 0.9998 \end{array} \right). \end{equation} One obvious defect of the above example is that $\lambda$ and the Cabibbo angle ${\bf V}_{12}$ are too big, as discussed in section~\ref{secpq}. Another is the degeneracy of the $s$ and $d$ quarks. Again, this is a generic feature of the examples based on the texture~(\ref{eqYukneg}). The reason is that the dominant terms in (\ref{eqYukneg}) are on the antidiagonal; the terms in the upper left triangle are orders of magnitude smaller. Therefore, (\ref{eqYukneg}) can be thought of as an antidiagonal matrix with a very small perturbation. Since a permutation of columns can cast it in an almost diagonal form, such a matrix will not lead to any appreciable flavor mixing {\em unless\/} two of the eigenvalues of $\bf Y\,Y^\dagger$ are degenerate. In such a case the choice of the basis in the eigenspace is arbitrary, and it is easy to obtain large mixing angles. We were looking for ${\bf V}_{12} \simeq 0.22$, so it is understandable for all the examples to have degenerate quark masses. The last thing we noticed is that the ``small'' entries (1,3), (3,1), (2,3), (3,2) of the CKM matrix~(\ref{eqV}) are much smaller than what we know from the experiment (even taking renormalization into account). The small mixing in the heavy flavor sector can be understood by noting that the heavy quarks are not degenerate in mass, and the magnitude of the mixing in this case is determined by the magnitude of the off-diagonal perturbation. \subsection{Allowing a Texture Factor} \label{sectf} Until now, we have been trying to keep all the factors of order unity in the Yukawa mass matrices equal to one. We tried to avoid the problem of singular mass matrices by restricting our search to matrices with enough texture zeros to be nonsingular. In this section, we want to make the matrices non-singular by introducing coefficients different from unity. To avoid introducing 27 free parameters, we make a rather arbitrary choice of the coefficients: we introduce one parameter, the ``texture factor'' (TF), which will multiply the (2,3), (3,2) and (3,3) elements of ${\bf Y}_u$ and ${\bf Y}_e$. The corresponding elements of ${\bf Y}_d$ are divided by TF. That is the minimal intervention needed to make the determinants nonzero in most cases. We chose to modify the entries in the heavy quark sector so that the predicted Cabibbo angle would not depend strongly on TF. We decided to divide, rather than multiply, in ${\bf Y}_d$, to make the determinant relations such as~(\ref{eqdet1}) minimally sensitive to TF. Finally, we considered only nine discrete values of TF: $-1$, $\pm2$, $\pm\sqrt{2}$, $\pm1/2$, $\pm1/\sqrt{2}$. With negative as well as positive charges allowed, but without requiring any particular pattern of texture zeroes, we were able to find better examples than before. With $g_s=0.7$ and the $X$ charges $$ \begin{tabular}{c|rrrrr} $i$ & $q_{Qi}$ & $q_{ui}$ & $q_{di}$ & $q_{Li}$ & $q_{ei}$ \\ \hline 1 & 5 & 3 & 2 & 6 & 5 \\ 2 & 4 & 2 & $-1$ & $-2$ & 4 \\ 3 & $-2$ & 1 & $-1$ & $-5$ & $-1$ \end{tabular} $$ we have $C_3=10$ and $\lambda=0.30$. With the texture factor ${\rm TF}=+2$, we find the fermion mass ratios \pagebreak[3] \begin{eqnarray*} \rule[-2pt]{0pt}{14pt} {m_u \over m_t} = 7.5\times10^{-6} \;, \qquad\quad {m_d \over m_b} &=& 3.3\times10^{-3} \;, \qquad\quad {m_e \over m_\tau} = 2.3\times10^{-3} \;,\\ \rule{0pt}{18pt} {m_c \over m_t} = 2.3\times10^{-3} \;, \qquad\quad {m_s \over m_b} &=& 3.1\times10^{-2} \;, \qquad\quad {m_\mu \over m_\tau} = 8.7\times10^{-2} \;,\\ \rule[2pt]{0pt}{16pt} {m_t \over f_u} = 2.0 \;, \qquad\qquad\qquad\; {m_b \over f_d} &=& 1.0 \;, \qquad\qquad\qquad\; {m_\tau \over f_d} = 1.0 \;, \end{eqnarray*} and the CKM matrix \begin{equation} \label{eqV2} {\bf V} = \left( \begin{array}{ccc} 0.98 & 0.20 & 5.0\times10^{-5} \\ -0.20 & 0.98 & 3.5\times10^{-4} \\ 2.1\times10^{-5} & -3.6\times10^{-4} & 1.0 \end{array} \right). \end{equation} Although the small elements of the CKM matrix are still about an order of magnitude too small, the above example has no other obvious defects. It reproduces the mass ratios~(\ref{eqratios}) and the Cabibbo angle fairly well, the bottom quark and the tau have equal masses at the unification scale, and the top quark is the heaviest. An even better fit (including the small elements of the CKM matrix) can be obtained for Kac-Moody level $k_{GUT}=2$, thanks to a much better agreement between (\ref{eqc318}) and (\ref{eqc3l}). This shows that, finally, it is possible to obtain realistic masses and mixings as a result of an anomalous $U(1)_X$ family symmetry. \section{Conclusions} \label{secconclusions} In this paper, we have examined the idea that the hierarchy parameter in Froggatt-Nielsen type models is related to a spontaneously broken anomalous $U(1)_X$ gauge symmetry left over from string compactification. Given the details of a string compactification, the condition of preserving supersymmetry (\ref{eqsqrt}) predicts the value of $\lambda$, hence the complete mass matrices, fermion masses and mixings. Without a complete string model at hand, we looked for model-indepen\-dent features. We found a strong constraint from anomaly cancellation (\ref{eqcratio}) that gives $\lambda$ in terms of the quark $X$ charges only. There is another constraint on those charges (\ref{eqc318}) from the known value of the product of all quark masses. Assuming $\lambda=0.22$, the two independent predictions (\ref{eqc3l}) and (\ref{eqc39}) agree as an order of magnitude relation. Their agreement in a definite model depends on $f_\theta$, the coefficient that enters $\lambda = f_\theta\langle\theta\rangle/M$, and on the Kac-Moody level of the gauge group, $k_{GUT}$. We also wanted to verify that it is possible to obtain realistic masses and mixings in this framework. We examine integer charge assignments satisfying the anomaly cancellation constraints, and, with the $\lambda$ determined by those charges, find some promising examples. It would be very interesting to do the same with a definite string model. The values for $f_\theta$, $k_{GUT}$ and the order unity texture factors in mass matrices would then be specified, leaving no free parameters. The method presented in this paper gives us a powerful tool to narrow down the set of possible string compactifications. We think that there will be only a small number (if any) of models compatible with the idea of predicting $\lambda$ from $U(1)_X$ charges. One desirable feature of the CKM matrix we have not addressed in this paper is CP violation. It is possible to introduce CP violation by assuming that the order unity texture factors are complex \cite{fn}, but that generically leads to large CP violation. Another way is to use a model with two $U(1)_X$ breaking fields, $\theta_1$ and $\theta_2$. The phase of the VEV of a single complex $\theta$ field can always be rotated away by a gauge transformation, but for two complex fields there is a gauge-invariant phase difference, so that in general we can make only one of them (say $\langle\theta_1\rangle$) real. The powers of $\langle\theta_1\rangle$ and $\langle\theta_2\rangle$ will give imaginary parts to the mass matrix elements. If $\langle\theta_2\rangle \ll \langle\theta_1\rangle$, then the imaginary parts will be necessarily small, leading to naturally small CP violation. Work on the details of the two-theta model is in progress. \section*{Acknowledgements} We have the pleasure to thank Jeff Harvey for suggesting the idea of this paper, and for many useful discussions until its completion. We would also like to acknowledge conversations with Tony Gherghetta, Aaron Grant, Erich Poppitz and Mihir Worah. This research was supported in part by NSF Grant No.\ PHY-9123780.
\section{Introduction} Magnetic fields are universally present in astronomical bodies ranging from the Earth to the distant quasars, but it is still unknown if magnetic fields permeate the universe as a whole. Astrophysical magnetic fields may arise due to the existence of a primordial magnetic field that grows as galaxies form. The discovery of an extragalactic magnetic field on scales larger than virialized systems (i.e., larger than clusters of galaxies) would reveal the presence of a primordial field. The existence of such a primordial field would help understand the origin of astrophysical magnetic fields and may open a new window into processes ocurring in the early universe. In this letter, we show that the study of ultra high energy cosmic rays (UHE CRs) with energies above $\simeq10^{18}{\,\rm eV}$ can probe primordial fields below the current upper bound. Our proposed method is complementary to traditional ones (e.g., Kronberg 1994) and more recent suggestions (e.g., Plaga 1995). At present, the most stringent constraint on large scale extragalactic fields comes from limits on the Faraday rotation of light coming to us from distant quasars. The upper bound on a widespread, all-pervading field is $\sim 10^{-9}$ Gauss (e.g., Kronberg 1994). There are weaker constraints derived from the synchrotron emission from nearby galaxy clusters (Kim et al.~1989) and the cosmic microwave background isotropy. UHE CR nucleons from extragalactic sources are attenuated in energy while propagating through the cosmic microwave background (CMB). Above a few $10^{19}{\,\rm eV}$, nucleons produce pions on the CMB photons and the energy of the cosmic ray nucleons is degraded rapidly, which is known as the Greisen-Zatsepin-Kuz'min (GZK) effect (Greisen 1966; Zatsepin \& Kuz'min 1966). Produced pions, on the other hand, decay into secondary leptons such as electrons, muons, and neutrinos, and photons. Since muons decay to electrons and neutrinos, the final secondary particles are photons, electrons, and neutrinos, among which photons are more readily detectable. Very energetic secondary photons and electrons couple to form an electromagnetic (EM) cascade. The $\gamma$-rays produce electron pairs on the CMB and the universal radio background photons (see Lee \& Sigl 1995, Lee 1995 and references therein for more detailed discussions). The resulting electrons (or positrons) in turn upscatter background photons via inverse Compton scattering (ICS), thus completing a cycle. It is through these two processes that an EM cascade develops in the intergalactic medium. As a result, if one propagates a purely protonic spectrum, one gets a processed nucleon spectrum with the GZK cutoff, and a secondary EM (photons and electrons) spectrum. The extragalactic magnetic field (EGMF) influences the UHE CR flux mainly through their charged components, namely the primary hadrons and the secondary electrons. In the energy range under consideration the hadrons are deflected with negligible energy loss due to synchrotron radiation whereas the electrons are negligibly deflected before they lose most of their energy. In \S 2, we explore the effect of the electron energy loss due to the EGMF on the secondary EM cascade spectrum. We then discuss the deflection of the hadronic component in the EGMF, in \S 3. Finally in \S 4, we summarize our findings. \section{Extragalactic Magnetic Field and Ultrahigh Energy {$\gamma$}-Ray Flux} The EGMF plays a crucial role in the development of the EM cascade. In the presence of the magnetic field, electrons lose energy via synchrotron radiation loss, which is given by \begin{equation} \frac{dE}{dt} = - \frac{e^4 B^2}{24\pi^2 m_e^4} E^2 = -\frac{2}{3} r_0^2 B^2 \left( \frac{E}{m_e} \right)^2, \end{equation} where $B$ is the strength of the large scale EGMF, $r_0$ is the classical electron radius, and $m_e$ is the electron mass. In fig.~1, we show the rates of ICS and synchrotron loss for electrons. Whereas the rate of ICS responsible for EM cascade development decreases with energy, the synchrotron loss rate increases with energy. Therefore, in a narrow energy range a transition occurs between a regime where ICS is dominant and electrons couple to photons efficiently and another where electrons are rapidly lost due to synchrotron loss and the cascade is suppressed. Below $\simeq 10^{20}{\,\rm eV}$ (the threshold for pair production on the radio background), cascade development in the absence of an EGMF would give rise to a generic power law photon spectrum with index $\simeq-1.5$. The above mentioned transition in the secondary {$\gamma$}-ray spectrum will therefore occur between this generic cascade shape and a synchrotron loss dominated spectrum. As long as synchrotron quanta can be neglected the latter is given by the photons produced ``directly'' by source injection or from pion production by nucleons, before undergoing pair production in the low energy photon background. In a magnetic field of strength $B$ measured in G (Gauss) the synchrotron spectrum produced by an electron of energy $E_e$ peaks at \begin{equation} E_{\rm syn} \simeq 6.8 \times 10^{13} \left( \frac{E_e}{10^{21} {\,\rm eV}} \right)^2 \left( \frac{B}{10^{-9}\,{\rm G}} \right) {\,\rm eV}\,,\label{synch} \end{equation} and falls off exponentially at higher energies. In the following we assume that the observable EM flux is energetically dominated by {$\gamma$}-rays and electrons with energy $E\la10^{21}{\,\rm eV}$. Then, according to eq.~[\ref{synch}], the contribution of synchrotron radiation to the {$\gamma$}-ray flux above $10^{18}{\,\rm eV}$ can be safely neglected as long as $B\la10^{-6}\,{\rm G}$ everywhere. The energy where the transition in the {$\gamma$}-ray spectrum occurs is in general a function of the magnetic field strength and the background photon spectrum, but {\em not\/} a function of the source distance or the injection spectrum. If we consider only the CMB, the relation between the transition energy $E_{\rm tr}$ and the magnetic field strength is $E_{\rm tr}\propto B^{-1}$. In order to include the less well known diffuse extragalactic radio background into consideration we adopt its usual description by a power law with an overall amplitude and a lower frequency cutoff as parameters (Clark, Brown, \& Alexander 1970) the latter one being the main source of uncertainty. Using a cutoff at 2 MHz as suggested by Clark et al.~(1970) the above relation is modified to \begin{equation} E_{\rm tr} \simeq 10^{19} \left(\frac{B}{10^{-9}\,{\rm G}}\right)^{-1.3}~{\,\rm eV} ~~(B \ga 10^{-10}\,{\rm G}) \ . \end{equation} For the same magnetic field a cutoff at lower frequencies would increase the rate of ICS of the then more abundant low frequency radio photons and thus the value for $E_{\rm tr}$ (see fig.~1). Assuming that the radio cutoff frequency lies somewhere in the range between 0.5 MHz and 3 MHz, for a given $E_{\rm tr}$ the EGMF strength $B$ is uncertain within about a factor 5. Therefore, for $B\ga10^{-10}\,{\rm G}$ it is possible to approximately determine the EGMF strength by searching for a dip in the $\gamma$-ray flux below $10^{20}{\,\rm eV}$ which would mark a transition between an ICS and a synchrotron loss dominated regime. For $B \la 10^{-10}\,{\rm G}$, this transition occurs above the pair production threshold on the radio background where the $\gamma$-ray flux increasingly depends on several unknown factors such as the charged cosmic ray flux above the GZK cutoff. Thus, even though the $\gamma$-ray flux can be comparable to the nucleon flux above $10^{20}{\,\rm eV}$, a discussion of possible magnetic field signatures in its spectrum would presently be too speculative. We developed a numerical code for the propagation of nucleons, photons, and electrons through the intergalactic medium which employs a transport equation formalism, the details of which can be found in Lee (1995). The observed UHE CR flux below $10^{20}{\,\rm eV}$ (see, e.g., Bird et al.~1994; Yoshida et al.~1995) is reproduced quite well by a diffuse distribution of sources injecting protons with a spectrum $\propto E^{-2.3}$ up to some maximal energy considerably beyond the GZK cutoff (Yoshida \& Teshima 1993; Sigl et al.~1995). For the calculations presented here we therefore adopted this proton injection spectrum with a maximal energy of $10^{22}{\,\rm eV}$ (see figures). The EGMF enters the calculation via the synchrotron loss of electrons. The transition between ICS and synchrotron loss domination can be easily seen in fig.~2, which shows the processed nucleon and photon spectra for a single source at a distance of 30 Mpc for a range of EGMF strengths. In general, one expects a distribution of cosmic ray sources rather than a single source at a fixed distance. In fig.~3, we show the diffuse spectrum from a continuous source distribution extending up to 1 Gpc. We assume a flat universe with zero cosmological constant and a Hubble constant of $H_0 = 75$ km sec$^{-1}$Mpc$^{-1}$, and a comoving source density scaling as $(1+z)^2$ in redshift $z$ as in some ``bright phase'' models of CR sources (e.g., Yoshida \& Teshima 1993; Hill \& Schramm 1985 and references therein). The results are not very sensitive to these choices. In the diffuse case, the {$\gamma$}-ray to nucleon flux ratio tends to be smaller than for a single source, and the spectral features are not as pronounced, but still detectable for $B \ga10^{-10}\,{\rm G}$. The ``extragalactic magnetic field'' in this analysis refers to the average component of the EGMF normal to the line of propagation. Primordial magnetic fields are expected to have very little structure on scales below $\sim $ few Mpc (Jedamzik, Katalinic, \& Olinto 1995), but condensed structures such as galaxies and clusters of galaxies can ``polute" the intergalactic medium with stronger magnetic fields on smaller scales. Fortunately, the effect of the EGMF on the $\gamma$-ray spectral shape discussed here is most sensitive to the average field on large scales. When the EM cascade goes through a strong field region, the electrons in the cascade lose energy rapidly and drop out of the UHE range. In contrast, the UHE nucleon and {$\gamma$}-ray fluxes are usually hardly affected directly by the radiation field of the intervening object (Stecker et al.~1991; Szabo \& Protheroe 1994; Norman, Melrose, \& Achterberg 1995). After escaping the object, the cascade redevelops and the cascade spectrum recovers quickly at only a slightly smaller amplitude. The influence of intervening objects decreases with increasing strength of the large scale field and becomes important only when the objects are very close to the observer (e.g., less than $\sim 5\,{\rm Mpc}$ away), or when their linear size is significant compared to the total propagation distance. We can derive conditions for the filling factors for such objects by requiring that they are much more sparsely populated along the line of sight to the source than the typical cascade regeneration length $s_c$ (e.g., $\sim$ 5 Mpc). For an average linear size $\bar l_{\rm iv}$ of the intervening objects their filling factor $f_{\rm iv}$ must satisfy \begin{equation} f_{\rm iv} \ll \frac{\bar l_{\rm iv}}{s_c}. \end{equation} For field galaxies the above relation is $f_g \ll 10^{-3}$, and for clusters of galaxies $f_c \ll 0.1$. The actual filling factors for galaxies, $f_g \la 10^{-7}$, and galaxy clusters, $f_c \la 10^{-4}$ (Kolb \& Turner 1992; Nichol, Briel, \& Henry 1994) satisfy these constraints. Therefore, intervening objects do not modify the above discussion substantially. One interesting exception may be the effect of nearby large structures such as the Virgo Cluster. If Virgo has strong magnetic fields, e.g.~of order $10^{-7}\,{\rm G}$ on Mpc scales, the UHE {$\gamma$}-ray flux from background sources might be modified across Virgo's angular extension (see fig.~2). The ability of future detectors to study EGMF features crucially depends on the {$\gamma$}-ray to nucleon flux ratio. For nearby strong sources, the secondary {$\gamma$}-ray flux could be measurable with instruments which are sensitive to ratios down to $\simeq1\%$. This could be achieved by the proposed Pierre Auger Project (see e.g., Boratav et al.~1992), which would also allow an angular resolution of $\simeq1^{\circ}$. The case of a diffuse source distribution is more challenging; the $\gamma$-ray flux is typically smaller and the EGMF feature is less pronounced than for a single source at moderate distances (see figs.~2 \& 3), but the dip in the $\gamma$ spectrum may still be detectable. Thus, measuring the {$\gamma$}-ray flux between $\simeq10^{18}{\,\rm eV}$ and $\simeq10^{20}{\,\rm eV}$ has the potential to either detect or find strong evidence against an EGMF $\ga 10^{-10}\,{\rm G}$. \section{Charged Cosmic Ray Deflection by Extragalactic Magnetic Fields} Here, we discuss the influence of the EGMF on the charged UHE CR flux from discrete sources. We restrict ourselves to the case of small deflection angles (for the opposite limit see, e.g. Wdowczyk \& Wolfendale 1979; Berezinskii, Grigor'eva, \& Dogel' 1989). In this case, the energy spectrum of charged UHE CRs from a given source is not significantly altered as compared to a straight-line propagation. However, if the sources are strong enough to cause an anisotropy in the UHE CR flux, the directional correlation of ``hot spots" with possible sources will depend on the EGMF. The following discussion relates to this anisotropic component of the charged UHE CR flux. As in \S 2, let us assume that the large scale EGMF can be characterized by a typical field strength $B$ and a coherence length $l_c$. Furthermore, we assume for the moment that the source distance $r$ is smaller than the energy attenuation length $\lambda=E(dE/dr)^{-1}$ for a charged cosmic ray of energy $E$ which can then be treated as approximately constant throughout propagation. For nucleons, $\lambda\simeq10$ Mpc above the GZK cutoff (at $E\simeq6\times10^{19}{\,\rm eV}$), and $\lambda\simeq1$ Gpc much below the GZK cutoff. A more sophisticated analysis would require a Monte Carlo simulation of both UHE CR propagation and deflection. However, since data on both UHE CR and the EGMF are so sparse to date, we feel that a qualitative discussion of the principle effects is more appropriate at the moment. We now consider two cases. (i) The source distance is smaller than the coherence length, $r\la l_c$. Then, in vectorial notation, the deflection angle ${\bf\alpha}$ is given by \begin{equation} {\bf\alpha}=-{Ze\over E}\,{\bf r}\times{\bf B}= 5.3^{\circ}\,Z\left({E\over10^{20}{\,\rm eV}}\right)^{-1}\left({r\over10\,{\rm Mpc}} \right)\left({B\over10^{-9}\,{\rm G}}\right)\, \left({\bf\hat r}\times{\bf\hat B}\right)\,,\label{def1} \end{equation} where ${\bf r}$ is the radius vector pointing to the source, $Z$ is the charge of the UHE CR component, ${\bf\hat r}={\bf r}/ \vert{\bf r}\vert$, and ${\bf\hat B}={\bf B}/\vert{\bf B}\vert$. Thus, a correlation between the UHE CR flux of charge $Z$ and energy $E$ and source counterparts, systematically shifted by an angle ${\bf\alpha}$, would indicate that $l_c\ga r$ and for a known source distance $r$ would allow to measure the combination $B({\bf\hat r}\times{\bf\hat B})$. The characteristic $E$- and $Z$ dependence of ${\bf\alpha}$ would provide an additional test for the hypothesis that the deflection is caused by an EGMF. (ii) The source distance is considerably larger than the coherence length, $r\gg l_c$. In this case the deflection angle undergoes a diffusion process during propagation and the source shape in the UHE CR flux will be smeared out over a typical angle \begin{equation} \alpha_{\em rms}\simeq{2\over\pi}{ZeB\over E}\left(rl_c\right)^{1/2}= 1.1^{\circ}\,Z\left({E\over10^{20}{\,\rm eV}}\right)^{-1} \left({r\over10\,{\rm Mpc}}\right)^{1/2} \left({l_c\over1\,{\rm Mpc}}\right)^{1/2} \left({B\over10^{-9}\,{\rm G}}\right)\,.\label{def2} \end{equation} Therefore, if sources appear spread out in the UHE CR flux of charge $Z$ and energy $E$ by a typical angle $\alpha$, this would indicate that $l_c\la r$ and for a known source distance $r$ would allow to measure the combination $Bl_c^{1/2}$. If the source distance is larger than the energy attenuation length, $r\ga\lambda$, eqs.~[\ref{def1}] and [\ref{def2}] tend to overestimate $\alpha$. In fact, in the limit $r\gg\lambda$ the deflection angle $\alpha$ ``saturates" as a function of $r$ and for approximately energy independent $\lambda$, $r$ has to be substituted by $\lambda$ and $\lambda/2$ in eqs.~[\ref{def1}] and [\ref{def2}], respectively. Secondary {$\gamma$}-rays produced by the interactions of the charged UHE CRs are also expected to correlate with the sources. Due to their continuous production they will be smeared out over angles which are typically somewhat smaller than given in eqs.~[\ref{def1}] and [\ref{def2}]. Sources which can act as suitable probes for the EGMF via the effects discussed above have to obey the following conditions apart from producing a detectable anisotropic UHE CR flux component: Their apparent angular size should be smaller than the deflection angle $\alpha$. The same pertains to the apparent angular radius of a possible high magnetic field region around the source if it can cause deflections in excess of $\alpha$. For example, a $10^{-6}\,{\rm G}$ field over a scale $\ga100\,{\rm kpc}$ is possible in galaxy clusters (see, e.g., Kronberg 1994) and would completely bend around a $10^{20}{\,\rm eV}$ proton. However, as long as a detectable proton flux emerges from such an object and the above conditions are fulfilled, it could still be a suitable probe of the EGMF. Finally, there should be no intervening high magnetic field regions between source and observer which could cause bending by more than $\alpha$. For example, for $r\ga l_c$, $\lambda$, this corresponds to the condition $l_{\rm iv}B_{\rm iv}\la\left(\lambda l_c\right)^{1/2}B$ for linear scale $l_{\rm iv}$ and strength $B_{\rm iv}$ of the intervening field. This condition could well be satisfied along most lines of sight since known objects with high field regions like galaxies and galaxy clusters have a small filling factor $f\la10^{-5}$. In light of these conditions we believe that some of the nearby galaxy clusters and powerful field radio galaxies could well be suitable EGMF probes since they are expected to contribute significantly to the UHE CR flux (Rachen, Stanev, \& Biermann 1993). The accuracy to which the EGMF bending can be determined is limited by the (to date) unknown additional bending by the galactic magnetic field of strength $B_g$ and scale height $l_g$. Thus, according to eq.~[\ref{def1}], the sensitivity of deflection measurements of the EGMF is restricted to field parameters satisfying $Br\ga10^{-9}\,{\rm G}\,{\rm Mpc} \left(B_g/\mu{\rm G}\right)\left(l_g/300\,{\rm pc}\right)$ where the fudge factors are the parameter values usually assumed for the galactic magnetic field. \section{Conclusions} We discussed how composition, spectrum, and directional distribution of UHE CR above $\simeq10^{18}{\,\rm eV}$ can be used to gain information about the large scale (a few to tens of Mpc) EGMF. Spectral features in the $\gamma$-ray flux are sensitive to field strengths in the range $\simeq10^{-10}-10^{-9}\,{\rm G}$. In a similar range, correlations between an anisotropic charged UHE CR flux component and possible sources could provide independent information on the EGMF including its polarization. Both effects should yield consistent estimates for the EGMF strength. Strong discrete sources detected in UHE CRs by future instruments with an angular resolution of $1^{\circ}$ or better and a sensitivity to $\gamma$-ray to nucleon flux ratios of $1\%$ or smaller would provide the best conditions for detecting an EGMF in the range $\simeq 10^{-10}-10^{-9}\,{\rm G}$. Since these conditions are not unreasonable, UHE CRs have the potential to provide important information on properties and origin of the EGMF. \acknowledgments This work was supported by the DoE, NSF and NASA at the University of Chicago, by the DoE and by NASA through grant NAG5-2788 at Fermilab, and by Alexander-von-Humboldt Foundation. S.L. acknowledges the support of the POSCO Scholarship Foundation in Korea.
\chapter{Introduction} A feature of recent developments in superstring theory is the emerging importance for a variety of non-perturbative phenomena of extended object, or `$p$-brane', solutions of the classical string theory. In particular, these solutions are crucial for an understanding of the various conjectured duality symmetries of both the heterotic and type II superstrings (see [\PKT] for a recent review). It is customary to call a $p$-brane `electric' if it is the source for a $(p+1)$-form potential in the effective field theory Lagrangian and `magnetic' if it is the source for the dual $(D-p-3)$-form potential. The word `source' may need some explanation here: one first solves the source-free equations of motion of the effective field theory; it is necessary to introduce an actual, `fundamental', source only if the analytic continuation of the source-free solution meets with a (timelike) singularity. Otherwise, no source is needed, but here one can interpret the extended object solution as an effective source on length scales that are long compared to the size of the object's core. In a D-dimensional spacetime the magnetic dual of an electric $p$-brane is a $\tilde p$-brane, where $\tilde p$ is related to $p$ by [\Nep] $$ \tilde p = D-p-4\ . \eqn\onea $$ It follows that a $p$-brane can carry {\it both} electric and magnetic charge only if $$ D= 2p+4\ , \qquad p=0,1,2,\dots \eqn\oneb $$ The simplest case is $D=4$ for which there arises the possibility of particles carrying both electric and magnetic charge, i.e. dyons. The next simplest case is $D=6$ for which there exists the possibility of dyonic strings. In fact, one can find a self-dual string in $D=6$, which is intrinsically dyonic because the two-form potential to which it couples has a self-dual field strength [\DL]. Other dyonic D=6 strings, which break more than half the supersymmetry, have been discussed in [\DFKR]. Here we consider the next case: membranes in D=8. Specifically, we present dyonic membrane solutions of N=2 D=8 supergravity that break half the supersymmetry. During the writing up of this work a paper [\BBO] presenting analogous results for D=6 dyonic strings appeared, in which the possibility of D=8 dyonic membranes was also mentioned. The reason for considering N=2 D=8 supergravity [\SaSe] is that this is the unique supersymmetric field theory (with no more than second order field equations) for which the field content includes a third-rank antisymmetric tensor gauge field. It may also be considered as the effective field theory for the $T^2$-compactified type II superstring. The N=2 D=8 supergravity theory has an $Sl(3;R)\times Sl(2;R)$ symmetry of the equations of motion. This group acts linearly on the various field strength tensors and their duals. These include a four-form field-strength $F$, and its dual, which transform according to the $({\bf 1},{\bf 2})$ representation of $Sl(3;R)\times Sl(2;R)$. The discrete subgroup $Sl(3;Z)\times Sl(2;Z)$ was conjectured in [\HT] to extend to a U-duality of the D=8 type II superstring theory; this discrete group contains the T-duality group $SO(2,2;Z)\equiv [Sl(2;Z)\times Sl(2;Z)]/Z_2$, which in turn contains an $Sl(2;Z)$ subgroup of the $Sl(2;R)$ group acting on the four-form field strength. This follows from the facts that (i) all non-perturbative U-duality symmetries are contained in the $Sl(3;Z)$ subgroup and (ii) $Sl(3;Z)$ acts trivially on $F$. Thus, although the $Sl(2;Z)$ group acts on $F$ via a generalized electromagnetic duality, this group is nevertheless a {\it perturbative} T-duality in the string theory context. This is to be expected from the fact that in the context of the type II superstring the three-form potential $A$ is a Ramond-Ramond (RR) field. A similar group-theoretical argument was used in [\HT] to show that electric {\it and} magnetic RR charges of the D=4 type II superstring transform irreducibly under T-duality. As we now see, the same is true in D=8. It should be noted that here we are using the term `T-Duality', in the context of $T^2$-compactifications of type II superstrings, to mean the identification of vacua of the resulting D=8 type II superstring under the discrete $SO(2,2;Z)$ subgroup of the $SO(2,2)$ classical symmetry group of the compactified theory. The analogous $SO(2,18;Z)$ T-Duality group of the heterotic string includes transformations which take $R\rightarrow 1/R$, where $R$ is the radius of an $S^1$ factor of $T^2$. For the type II superstrings this $R\rightarrow 1/R$ transformation (also called T-Duality) interchanges the type IIA and type IIB superstrings, which are therefore equivalent to a single D=8 type II superstring. Such transformations are {\it not} realized as gauge symmetries of this D=8 theory, and therefore are not included in the $SO(2,2;Z)$ T-Duality group. There is a consistent truncation of the N=2 D=8 supergravity in which the only surviving fields are the spacetime metric, $g_{\mu\nu}$, a scalar, $\sigma$, a pseudoscalar $\rho$ and a three-form gauge potential, $A$, for which $F=dA$ is its four-form field-strength. The Lagrangian of this truncated theory is $$ \eqalign{ {\cal L} = N\Bigg\{\sqrt{-g}\big[ &R - 2\partial_\mu \sigma\partial^\mu\sigma - 2e^{4\sigma}\partial_\mu \rho\partial^\mu\rho -{1\over 12}e^{-2\sigma}F_{\alpha\beta\gamma\delta}F^{\alpha\beta\gamma\delta}\big] \cr &-{1\over 144}\varepsilon^{\mu\nu\rho\sigma\alpha\beta\gamma\delta}\rho F_{\mu\nu\rho\sigma}F_{\alpha\beta\gamma\delta} \Bigg\}\ ,} \eqn\onec $$ where $N$ is a normalization factor, which we can choose at our convenience. The coefficient of the $\varepsilon\rho FF$ term is crucial to the results to follow so we should point out that we disagree by a factor of three with the coefficient of this term given in [\SaSe]. The coefficient can be simply determined by dimensional reduction of the D=11 supergravity theory, which was the method used in [\SaSe], but this leads to the coefficient used here rather than that of [\SaSe]. The $\sigma$ and $\rho$ kinetic terms of \onec\ constitute a sigma model with target space $Sl(2;R)/U(1)$. It is convenient to introduce the complex field $$ \lambda = 2\rho + ie^{-2\sigma}\ , \eqn\comfield $$ taking values in the upper half complex plane, since the $Sl(2;R)$ group acts on $\lambda$ by fractional linear transformations. Since the asymptotic value of $\lambda$ is undetermined by the equations of motion, the possible vacua correspond to points in the upper half plane. However, T-duality of the type II D=8 superstring theory implies that points that lie in an orbit of an $Sl(2;Z)$ subgroup of $Sl(2;R)$ correspond to equivalent vacua. Thus, the moduli space of vacua in the string theory context is, assuming T-duality, the fundamental domain of $SL(2;Z)$ in the upper half complex plane. Note the similarity of the above discussion with that of S-duality in the heterotic string. The main difference, apart from the obvious one that here we are dealing with a four-form rather than a two-form field strength, is that the scalar field $\sigma$ in \onec\ is {\it not} the dilaton. In fact, the dilaton has been set to zero in the truncation leading to \onec. If $\sigma$ were the dilaton, the $Z_2$ subgroup of $Sl(2;Z)$ that exchanges $F$ with its dual and takes $\sigma$ to $-\sigma$ would be non-perturbative in the context of the type II string theory. As explained above, this is not the case. An alternative explanation is provided by string-string duality, as will shall see shortly. We shall be interested in infinite planar membrane solutions of the equations of motion of \onec\ that are asymptotically flat as one approaches spatial infinity in non-coplanar directions; we shall call this `transverse spatial infinity', which is topologically $S^4\times R^2$. Membrane solutions can be characterised by their electric and magnetic number densities $$ q = {N\over e}\oint\! G \qquad p= {e\over 2\pi}\oint\! F \eqn\chargetwo $$ where the integral is over a 4-sphere cross-section of transverse spatial infinity, $e$ is an arbitrary unit of `electric' charge, and the two-form $G$ is related to the Hodge dual $\tilde F$ of $F$ by $$ G\equiv e^{-2\sigma}\tilde F -2\rho F \ . \eqn\chargetwo $$ We shall require an asymptotic translational invariance in directions coplanar with the membrane so that these number densities are actually constant; we shall refer to these constants as the membrane `charges'. Their conservation follows from the fact that the combined equations of motion and Bianchi identities of the field-strength four-form $F$ can be written as $d{\cal F}=0$ where ${\cal F}$ is the $Sl(2;R)$ doublet $$ {\cal F} = (F,G)\ . \eqn\chargethree $$ We shall choose the constants $N$ and $e$ such that $$ q = {1\over \Omega_4}\oint G \qquad p= {1\over \Omega_4}\oint F \eqn\charge $$ where $\Omega_4=2\pi^2$ is the volume of the unit 4-sphere. With this choice, the charges $(p,q)$ form an $Sl(2;R)$ doublet. As shown in [\Nep], the electric and magnetic charges of extended objects are subject to a generalization of the Dirac quantization condition. However, just as the Dirac quantization condition must be replaced, in the context of dyons, by the Schwinger-Zwanziger quantization condition so, in the context of dyonic extended objects, the Nepomechie-Teitelboim (N-T) quantization condition must be replaced by an extended object analogue of the Schwinger-Zwanziger quantization condition. With the above choice of normalization constant, $N$, and electric charge unit, $e$, this generalized N-T quantization condition for two dyonic membranes with charges $(p,q)$ and $(p',q')$ takes the simple (manifestly $Sl(2;R)$ invariant) form $$ qp'-q'p \ \in \ Z\ . \eqn\Dizzy $$ As for dyons in D=4 [\Witb], this formula allows fractional $q$ for dyonic membranes, but the consequences for dyonic membranes are not quite the same as those for dyons because one cannot take for granted the existence of purely electric membranes in the quantum theory. In [\GHT] it was shown how an analogue of the Bogomolnyi-Gibbons-Hull bound for particle-like solutions of Maxwell/Einstein theory can be derived for $p$-brane solutions of certain antisymmetric tensor generalizations of Maxwell/Einstein theory. The precise interactions of the antisymmetric tensor field, e.g. the coefficient of possible Chern-Simons terms was crucial to this result. In all cases, the interactions were precisely those for which the bosonic field theory could be interpreted as a consistent truncation of a supergravity theory. Since this condition is satisfied by the Lagrangian \onec\ one would expect to be able to derive a similar bound on the tension of membrane solutions of its equations of motion; this case is not covered by the results of [\GHT] because Lagrangians with scalar fields were not considered there. This expectation is correct; we shall show that the tension, $M$, of membrane solutions of \onec\ satisfies the $Sl(2;R)$ invariant bound $$ M^2 \ge {1\over 4} \Big[ e^{2\langle\sigma\rangle}\big(q+2\langle\rho\rangle p\big)^2 + e^{-2\langle\sigma\rangle} p^2\Big] \ , \eqn\abog $$ where $\langle \rho\rangle$ and $\langle\sigma\rangle$ are the asymptotic values of $\rho$ and $\sigma$. Solutions which saturate the bound are `supersymmetric' in that they admit Killing spinors. The purely electric and magnetic D=8 supersymmetric membrane solutions, with $\rho\equiv0$, have been given previously [\DL]. The supersymmetric membrane solutions we construct here differ in that they have non-constant axion field and carry both electric and magnetic charge, i.e. they are `dyonic'. There is a U(1) parameter family of these solutions for each value of the asymptotic values of $\sigma$ and $\rho$, corresponding to the U(1) stability subgroup of $Sl(2;R)$ acting on the upper-half plane by fractional linear transformations. Although only a $Z_2$ family of these will survive quantization, the identification of vacua related by a transformation in the $Sl(2;Z)$ T-duality subgroup of $Sl(2;R)$ allows us to find $Sl(2;Z)$ orbits of membrane solutions about equivalent vacua, as has been done previously for particle-like solutions in D=4 [\STW,\KO]. Almost all such solutions are dyonic. One motivation for our work derives from a recently suggested D=8 membrane/membrane duality [\PKTc]. The point here is, firstly, that while the purely electric membrane solution of N=2 D=8 supergravity theory can be interpreted as the membrane solution of D=11 supergravity in a $T^3$ compactified spacetime, the purely magnetic one can be interpreted as a double dimension reduction of the fivebrane solution of D=11 supergravity\foot{This was stated in [\PKTc]; here we verify it.}. Secondly, the worldvolume action of this magnetic membrane is that of a D=11 supermembrane in a $T^3$ compactified spacetime (and not that of a D=8 supermembrane, as one might have guessed; the extra three coordinates come from the antisymmetric tensor in the fivebrane's worldvolume action). This suggests a complete non-perturbative equivalence between the electric and magnetic membranes. This equivalence would be guaranteed in string theory by non-perturbative T-duality. Unfortunately, this cannot be established in string perturbation theory, but one can reverse the logic and use the evidence of membrane/membrane duality given in [\PKTc] and the results presented here as evidence for the non-perturbative validity of T-duality. Another motivation comes from the conjectured non-perturbative equivalence of the $K_3\times T^2$ compactified type II superstring theory with the toroidally compactified heterotic string theory [\HT], i.e. the `string-string duality' for which there is now considerable evidence. Many recent papers dedicated to tests of this conjecture have taken as their starting point the related conjecture that the D=6 string theories obtained by compactification of the type IIA superstring on $K_3$ and the heterotic string on $T^4$ are non-perturbatively equivalent [\Witten]. Given this D=6 equivalence, the equivalence in D=4 follows upon further compactification on $T^2$. S-duality of the heterotic string [\FILQ,\Sen] can then be re-interpreted as T-duality of the type II superstring [\SS,\Duff,\Witten]. This approach to understanding D=4 S-duality via the heterotic/type II equivalence can be characterised by the motto ``10 to 6 and then to 4''. Our work can be viewed as a first step towards an understanding of heterotic S-duality via the alternative ``10 to 8 and then to 4'' approach. The first step is a $T^2$ compactification of both the type II and the heterotic string to D=8. A subsequent compactification of the D=8 type II superstring on $K_3$ yields a D=4 string theory which, according to string-string duality, is equivalent to the $T^4$ compactified D=8 heterotic string. The spectrum of this D=4 string theory includes dyons which arise, in the type II interpretation, as wrapping modes of D=8 dyonic membranes around the 22 fundamental homology 2-cycles of $K_3$. These dyons are charged with respect to the 22 D=4 two-form field strengths, $F^I\ (I=1,2,\dots,22)$, arising from the D=8 four-form field strength $F$ via the ansatz $$ F(x,y)= F^I(x)\wedge\omega_I(y) \ , \eqn\ansatz $$ where $\omega_I$ span the 22-dimensional space of harmonic two-forms on $K_3$. These D=4 dyons are non-perturbative RR states, even the purely electric ones; they form multiplets of the $Sl(2;Z)$ (type II) T-duality subgroup descending from the T-duality group in D=8. Note that the full type II T-duality group in D=4 is the same as the full type II T-duality group in D=8 because $K_3$ has no continuous isometries. According to string-string duality the type II RR dyons just discussed must appear in the spectrum of an equivalent heterotic string. Moreover, one expects the purely electric particles among them to appear as {\it perturbative} states in view of the generally accepted opinion that {\it all} purely electric states of the heterotic string are perturbative. Since their dyonic $Sl(2;Z)$ partners are necessarily non-perturbative, the $Sl(2;Z)$ group that relates them must then be a {\it non-perturbative} duality group of the heterotic string, i.e. the S-duality group. Thus, the existence of the $Sl(2;Z)$ multiplets of dyonic D=8 membranes provides further confirmation of the interchange of S and T duality effected by string-string duality. In the following, we begin with a presentation of the dyonic membrane solutions of the field equations of the Lagrangian \onec. We then explain how these solutions were found and why their tension saturates a Bogomolnyi-Gibbons-Hull type bound. We also exhibit the Killing spinors admitted by these solutions, thereby establishing their supersymmetry. We then discuss the global structure of the dyonic membranes and their interpretation as solutions of D=11 supergravity. We conclude with some further comments on the significance of our results. \chapter{D=8 dyonic membranes} The field equations of the Lagrangian \onec\ are $$ \eqalign{ G_{\mu\nu} &= 2T_{\mu\nu} \cr \partial_\mu \big(\sqrt{-g}\; e^{-2\sigma}F^{\mu\nu\rho\sigma}\big) &=- 2\big(\partial_\mu\rho\big) \tilde F^{\mu\nu\rho\sigma} \cr \partial_\mu\big(\sqrt{-g}\; e^{4\sigma}\partial^\mu\rho \big) &= {1\over 24} F_{\mu\nu\rho\sigma}\tilde F^{\mu\nu\rho\sigma} \cr \partial_\mu\big(\sqrt{-g}\; \partial^\mu\sigma\big) &= \sqrt{-g}\big[2 e^{4\sigma}(\partial\rho)^2 -{1\over24} e^{-2\sigma} F^2 \big]\ ,} \eqn\aonec $$ where $$ \eqalign{ T_{\mu\nu} = &\big[\partial_\mu\sigma\partial_\nu\sigma - {1\over2} g_{\mu\nu}(\partial\sigma)^2\big] + e^{4\sigma} \big[\partial_\mu\rho\partial_\nu\rho - {1\over2} g_{\mu\nu}(\partial\rho)^2\big]\cr &\qquad +{1\over6}e^{-2\sigma} \big[ F_{\mu\alpha\beta\gamma}F_{\nu}{}^{\alpha\beta\gamma} -{1\over8}g_{\mu\nu}F^2 \big] \ ,} \eqn\bonec $$ and $$ \tilde F^{\mu\nu\rho\sigma} \equiv {1\over 24} \varepsilon^{\mu\nu\rho\sigma\alpha\beta\gamma\delta} F_{\alpha\beta\gamma\delta}\ . \eqn\donec $$ We shall consider field configurations representing an infinite planar membrane and choose coordinates such that it is aligned with the $x^1\equiv y$ and $x^2\equiv z$ axes. We shall look for product metrics in which the metric of the five-dimensional `transverse' space is conformally flat and may therefore be parameterised by the coordinates ${\bf x}\equiv (x^3,\dots,x^7)$ of an associated five-dimensional Euclidean space, $\bb{E}^5$. There are certainly many solutions of the field equations \onec\ within this class of field configurations, but we shall concentrate on those that admit Killing spinors. We shall first present these solutions. Then, in the following section, we shall explain how they were obtained and why they are supersymmetric. We shall present the solutions in terms of the complex field $\lambda$ defined in \comfield. If we fix boundary conditions such that the spacetime is asymptotically flat as $|{\bf x}|\rightarrow \infty$, and such that $$ \lambda \rightarrow i\ , \eqn\bcs $$ then the following multi-membrane field configurations solve \aonec\ for arbitrary angular parameter $\xi$: $$ \eqalign{ ds^2 &= H^{-{1\over2}}[-dt^2 + dy^2 + dz^2] + H^{1\over2} d{\bf x}\cdot d{\bf x}\cr F &= {1\over2}\cos\xi\, (\star dH) + {1\over2}\sin\xi\, dH^{-1}\wedge dt\wedge dy\wedge dz \cr \lambda &= {\sin 2\xi\,(1-H) + 2iH^{1\over2}\over 2 (\sin^2\xi + H\cos^2\xi)}\ .} \eqn\dyons $$ Here, the symbol $\star$ indicates the Hodge dual in $\bb{E}^5$ and $$ H= 1 + \sum_{n=1}^N{\mu_n\over |{\bf x}-{\bf x}_n|^3} \eqn\harm $$ for $n$ arbitrary constants $\mu_n$ associated with the $N$ points ${\bf x}={\bf x}_n$, for any finite value of $N$. That is, $H({\bf x})$ solves the Laplace equation on $\bb{E}^5$ with an arbitrary number of point sources and is such that $H\rightarrow 1$ as $|{\bf x}|\rightarrow \infty$. The constants $\mu_n$ are proportional to the ADM tension of each membrane solution. Specifically, for a one membrane solution with parameter $\mu$ the ADM tension is $$ M= {3\over4}\mu\ . \eqn\tension $$ We have presented the solutions for a specially chosen asymptotic value of $\lambda$ because a solution with any other asymptotic value of $\lambda$ can be found by making use of the $Sl(2;R)$ invariance of the field equations. As stated earlier, this $Sl(2;R)$ group acts on $\lambda$ by fractional linear transformations: $$ \lambda \rightarrow {a\lambda + b\over c\lambda + d}\ , \eqn\onee $$ where $a,b,c,d$ are real numbers such that $ad-bc=1$. The $Sl(2;R)$ group acts on the four-form doublet ${\cal F} = (F, G)$ by a generalization of electromagnetic duality. Specifically, if $\lambda$ is transformed as in \onee, then the associated transformation of ${\cal F}$ is $$ {\cal F}\rightarrow (F, G) \pmatrix{d&-b\cr -c& a}\ . \eqn\onefb $$ Since there is a $U(1)$ isotropy subgroup of $Sl(2;R)$ that does not change the asymptotic value, $\langle\lambda\rangle$, of $\lambda$, there must be a $U(1)$ family of solutions for each choice of $\langle\lambda\rangle$. This is the significance of the angular parameter $\xi$ in \dyons. This $U(1)$ group is an analogue of the electromagnetic duality group since it takes a purely electric or purely magnetic solution into a dyonic one. Thus, the general solution of \dyons\ can be obtained by a $U(1)$ transformation of the purely magnetic solution $$ \eqalign{ ds^2 &= H^{-{1\over2}}[-dt^2 + dy^2 + dz^2] + H^{1\over2} d{\bf x}\cdot d{\bf x}\cr F &= {1\over2}\star dH \cr \lambda &= iH^{-{1\over2}} \ .} \eqn\dyonstwo $$ However, because of charge quantization, this classical $U(1)$ symmetry will be broken to $Z_2$ in the quantum theory; there will be some `preferred' value of $\langle\lambda\rangle$ for which only the purely electric or purely magnetic solutions survive (by analogy with D=4 dyons one might suppose that $\langle\lambda\rangle =i$ is the `preferred' value; we shall examine this hypothesis in more detail later). It might therefore appear that the more general dyonic membrane solutions of \dyons\ are irrelevant to the type II string theory, at least for the `preferred' value of $\langle\lambda\rangle$. However, the sigma-model target space of \onec\ is only required by supersymmetry to be {\it locally} isometric to the coset space $SL(2;R)/U(1)$. It may differ globally since it is possible to identify points on this space that differ by the action of $Sl(2;Z)$. Thus, the true sigma-model space could be $$ {\cal{M}} = Sl(2;Z)\backslash Sl(2;R)/U(1)\ . \eqn\true $$ In this case the true moduli space is not the entire upper-half $\lambda$-plane but rather the fundamental domain of $Sl(2;Z)$ in the upper half plane. In the context of the D=8 type II superstring theory, T-duality implies that this is indeed the true moduli space of vacua, so vacua which differ by the action of $Sl(2;Z)$ should be identified. Thus an $Sl(2;Z)$ transformation of the purely magnetic membrane solution \dyonstwo\ will produce a new solution with a different, but {\it equivalent}, value of $\lambda$, and this solution will have an effective non-zero value of $\xi$, i.e. it will be dyonic. Actually, we shall find a more general class of dyonic solutions by applying this procedure to the dyonic solutions \dyons\ rather than to the purely magnetic solution \dyonstwo, i.e. we allow for an arbitrary initial value of the angular parameter $\xi$. First we make an $Sl(2;R)$ transform of the solution \dyons\ to arrive at $$ \eqalign{ ds^2 &= H^{-{1\over2}}[-dt^2 + dy^2 + dz^2] + H^{1\over2} d{\bf x}\cdot d{\bf x}\cr F &= {1\over2}e^{2\langle\sigma\rangle}\Big( \cos\psi \star dH + \sin\psi\; dH^{-1} \wedge dt\wedge dy\wedge dz\Big) \cr \lambda &= 2\langle \rho\rangle + e^{-2\langle\sigma\rangle} \cdot {(1-H)\sin 2\psi + 2i H^{1\over2}\over 2(H\cos^2\psi + \sin^2\psi)} \ ,} \eqn\dyonsthree $$ where $$ e^{-2\langle\sigma\rangle} = {1\over c^2 + d^2}\ ,\qquad 2\langle\rho\rangle = {bd + ac\over c^2 + d^2}\ , \eqn\dyonsfour $$ and the new angular parameter $\psi$ is given by $$ \tan\psi = {d\sin\xi + c\cos\xi\over d\cos\xi -c\sin\xi}\ . \eqn\adyonsfour $$ Then, we restrict $a,b,c,d$ to be integers to obtain the dyon solutions with $\langle\lambda\rangle \cong i$. By construction, these solutions form a representation of $Sl(2;Z)$. Note that the set of dyon solutions obtained in this way will contain a purely magnetic solution if and only if $\tan\xi$ is rational. If this condition is satisfied then there will also be a purely electric solution. Clearly, a similar set of dyonic membrane solutions can be found for any other initial choice of $\langle\lambda\rangle$. However, if initially $\langle\lambda\rangle \ne i$, then the $Sl(2;Z)$ subgroup is not found by simply restricting $a,b,c,d$ to be integers. Rather, the elements of the $Sl(2;Z)$ subgroup are similarity transforms of matrices with integer entries. \chapter{Killing spinors and the Bogomol'nyi Bound} We have claimed that the dyonic membrane solutions presented above are supersymmetric, i.e. that they admit Killing spinors. We shall now elaborate on this point. A Killing spinor is a spinor field, $\epsilon$, that is in the kernel of a first-order Lorentz-covariant Dirac-type operator $\hat{\cal D}$, i.e. $\hat{\cal D}\epsilon=0$, where a minimal condition on $\hat{\cal D}$ is that the vector field $\bar\epsilon\gamma^\mu\epsilon$ is Killing if $\epsilon$ is. In the context of field theories with scalar and vector fields, this condition limits, but does not define, $\hat{\cal D}$. Within the context of a supergravity theory, $\hat{\cal D}$ is defined by the gravitini transformation laws, but an alternative intrinsic definition is possible in the context of an {\sl a priori} arbitrary bosonic Lagrangian via the modified Nester tensor $$ \hat E^{\mu\nu} = {1\over2}\bar\epsilon\Gamma^{\mu\nu\rho}\hat{\cal D}_\rho\epsilon + c.c.\ . \eqn\Nester $$ This is because the operator $\hat{\cal D}$ is fixed, if it exists, by the requirement that $$ {\cal D}_\nu\hat E^{\mu\nu} = \overline{\hat{\cal D}_\nu\epsilon}\; \Gamma^{\mu\nu\rho}\hat{\cal D}_\rho\epsilon - {1\over2} \bar\chi \Gamma^\mu\chi\ , \eqn\crucial $$ as a consequence of the field equations, for some complex spinor $\chi$. This requirement also fixes $\chi$. The significance of the relation \crucial\ is that it allows the derivation of a bound on the mass per unit $p$-volume, i.e. the tension, of configurations that are subject only to the boundary conditions at transverse spatial infinity satisfied by $p$-brane solutions of the equations of motion [\GHT]. It can happen that the field equations of a given Lagrangian are such that \crucial\ is not satisfied by any operator $\hat{\cal D}$ for any spinor $\chi$. In this case a bound on the tension cannot be derived by this method. Conversely, requiring that such a bound be derivable in a Lagrangian whose interactions are parameterised by arbitrary functions of the scalar fields can fix these functions. For example, allowing arbitrary interactions of $\sigma$ consistent with the requirement that the field equations be of second order, and an arbitrary coefficient of the $\rho F\tilde F$ term, one finds that the only Lagrangian in this class for which an energy bound on the membrane tension can be derived is precisely the Lagrangian of \onec. For the case in hand, one finds that $$ \hat{\cal D}_\mu \epsilon \equiv {\cal D}_\mu\epsilon -{1\over2} \gamma_9\epsilon\; e^{2\sigma} \partial_\mu \rho + {1\over 96}\Gamma^{\alpha\beta\gamma\delta} \Gamma_\mu\epsilon\; e^{-\sigma} F_{\alpha\beta\gamma\delta}\ , \eqn\oneh $$ and $$ \chi = \Gamma^\mu\epsilon\; \partial_\mu\sigma - \gamma_9\Gamma^\mu\epsilon\; e^{2\sigma}\partial_\mu\rho -{1\over 48} \Gamma^{\alpha\beta\gamma\delta}\epsilon\; e^{-\sigma}F_{\alpha\beta\gamma\delta}\ . \eqn\onei $$ The matrix $\gamma_9$ is defined by $$ \gamma_9 =\Gamma^{\underline 0}\Gamma^{\underline 1}\cdots\Gamma^{\underline 7} \eqn\onej $$ where the underlining indicates a flat space Dirac matrix. It follows from \oneh\ that $$ \hat E^{\mu\nu} = E^{\mu\nu} -{1\over2} e^{2\sigma} (\bar\epsilon \Gamma^{\mu\nu\alpha}\gamma_9 \epsilon)\partial_\alpha \rho - {1\over4} e^{-\sigma}\bar\epsilon(F^{\mu\nu\alpha\beta}\Gamma_{\alpha\beta} -\tilde F{}^{\mu\nu\alpha\beta}\Gamma_{\alpha\beta}\gamma_9)\epsilon \eqn\aonej $$ where $E^{\mu\nu}$ is the standard Nester tensor. Note that the Dirac conjugate $\bar\psi$ of a spinor $\psi$ is defined by $$ \bar\psi = \psi^{\dagger} \Gamma^{\underline 0}\ , \eqn\conjugate $$ so that $\bar\psi\Gamma^{\underline 0}\psi$ is negative definite. Note also that the Lorentz invariant $\bar\psi\psi$ is pure imaginary (for commuting spinors) since $\Gamma^{\underline 0}$ is anti-Hermitian\foot{In the Majorana basis, in which the matrices $\Gamma^\mu$ are (for D=8) pure imaginary, $\Gamma^{\underline 0}$ is symmetric and equal to $i$ times the charge conjugation matrix (which is symmetric for D=8).} As explained in the introduction, the relevant concept for defining membrane charges is transverse spatial infinity, which has topology $S^4\times R^2$. It is convenient to choose periodic boundary conditions to convert this to $S^4\times T^2$, i.e. we consider the membrane to be wrapped around a large two-torus. The energy per unit area, $M$, is then the, now finite, total energy divided by the volume, $V_2$, of the two-torus. This energy can expressed as an integral over the $S^4\times T^2$ surface at spatial infinity. Specifically, if ${\bf P}$ is the total transverse 5-momentum per unit area, such that $M=\sqrt{-|{\bf P}|^2}$, then [\GHT] $$ \bar\epsilon_\infty {\bf \Gamma}\cdot {\bf P}\epsilon_\infty = {1\over 2V_2 \Omega_4} \oint_\infty \! dS_{\mu\nu} E^{\mu\nu}\ , \eqn\boga $$ where $\Omega_4$ is the volume of the unit 4-sphere. With appropriate asymptotic fall off conditions on the metric, and assuming that $$ \epsilon \rightarrow \epsilon_\infty \eqn\eplim $$ as $|{\bf x}|\rightarrow \infty$, for some constant spinor $\epsilon_\infty$, \boga\ can be rewritten as $$ \bar\epsilon_\infty {\bf \Gamma}\cdot {\bf P}\epsilon_\infty = {1\over 2 \Omega_4} \oint_\infty \! dS_{ij} E^{ij}\ , \eqn\bogb $$ where the integral is now over the 4-sphere at spatial infinity and the index $i$ is associated with the coordinates ${\bf x}$ of the transverse space. Assuming that the only components of $F$ that are non-vanishing at transverse spatial infinity are $F_{ijkl}$ and $F_{tyzi}$, and that these components depend asymptotically only on $x^i$, one has that $$ \eqalign{ {1\over 2V_2\Omega_4} \oint_\infty \! dS_{\mu\nu}\hat E^{\mu\nu} &= {1\over 2 \Omega_4} \oint_\infty \! dS_{ij} \hat E^{ij}\cr &= \bar\epsilon_\infty \Big[{\bf \Gamma}\cdot {\bf P} - {1\over 8 \Omega_4} e^{-\langle\sigma\rangle} \Gamma_{kl}\oint_\infty \! dS_{ij}\Big( F^{ijkl} - \tilde F^{ijkl}\gamma_9\Big)\Big]\epsilon_\infty\ , } \eqn\abog $$ since the $\partial\rho$ term in \aonej\ does not contribute to the integral. {}From the definitions \charge\ of the charges $(p,q)$ one then finds that $$ {1\over 2V_2\Omega_4} \oint_\infty \! dS_{\mu\nu}\hat E^{\mu\nu} = \bar\epsilon_\infty K \epsilon_\infty \eqn\bogc $$ where $$ K= {\bf \Gamma}\cdot {\bf P} -{1\over2} \Big[ e^{\langle\sigma\rangle}(q+2\langle\rho\rangle p) \Gamma_{yz} - e^{-\langle\sigma\rangle} p\; \Gamma_{yz}\gamma_9\Big]\ . \eqn\abogc $$ Using Gauss's law, the relation \crucial, and choosing $\epsilon$ to satisfy a `modified Witten condition', one can prove that the integral on the left hand side of \bogc\ is positive semi-definite, subject to the usual assumptions. It follows that the Dirac matrix $K$ is positive semi-definite, which implies the bound \abog\ quoted in the introduction. This bound is saturated by solutions of the equations of motion for which there exists a spinor $\epsilon$ such that $$ \hat{\cal D}_\mu\epsilon=0\ , \qquad\qquad \chi=0\ . \eqn\Killing $$ Non-trivial solutions of these relations, i.e. those for which $M\ne0$, require $\epsilon$ to satisfy a condition of the form $$ \big[\alpha({\bf x}) \Gamma_* + \beta({\bf x})\Gamma_*\gamma_9\big]\epsilon({\bf x}) =\epsilon ({\bf x})\ , \eqn\constraint $$ where $$ \Gamma_* = \Gamma^{\underline 0}\Gamma^{\underline 1}\Gamma^{\underline 2}\ . \eqn\onek $$ and $\alpha$ and $\beta$ are functions such that $$ \alpha^2 + \beta^2=1 \ . \eqn\aonek $$ This can be seen from the fact that the spinor $\epsilon$ must be an eigen-spinor of the matrix $K$ with zero eigenvalue. The angular parameter $\xi$ enters into the solutions \dyons\ as the limit of the ratio of the functions $\alpha$ and $\beta$, i.e. $$ \lim_{|{\bf x}|\rightarrow \infty} \Big({\alpha\over\beta}\Big) = \tan\xi\ . \eqn\bonek $$ The multi-dyon solutions \dyons\ were obtained by substituting an appropriate ansatz into the relations \Killing. The constraint \constraint\ reduces the dimension of the space of Killing spinors to half that of the constant Killing spinors of the vacuum. Thus, the solutions we find in this way will break half the supersymmetry. Furthermore, they saturate the bound \abog\ by construction, so their membrane tension is given by the formula $$ M^2 = {1\over 4} \Big[ e^{2\langle\sigma\rangle}(q+2\langle\rho\rangle p)^2 + e^{-2\langle\sigma\rangle} p^2\Big] \ . \eqn\bogi $$ where $M$ is related to the constants $\mu_n$ appearing in the solutions by $$ M= {3\over4} \sum_{n=1}^N \mu_n\ . \eqn\bogsum $$ This bound does not imply that each constant $\mu$ is individually positive but it is easy to see that this must in fact be the case. The point is that \bogsum\ is independent of the positions of the membranes, so that we may consider a limit in which they become arbitrarily far apart. In this limit we may also take the infinite radius limit of a four-sphere surrounding any given membrane of tension $\mu$. This four-sphere approaches transverse spatial infinity in this limit but encloses only the chosen membrane, so that $\mu$ must be positive. Because of this, the only singularities of the metric are at the `centres' ${\bf x}={\bf x}_n$. The question whether these are real singularities or merely coordinate singularities will be addressed in the following section. From \onefb\ we see that the $Sl(2;R)$ transformation of (p,q) is $$ (p,q)\rightarrow (p,q)\pmatrix{d&-b\cr -c& a}\ . \eqn\abogsum $$ Given that $\langle\sigma\rangle$ and $\langle\rho\rangle$ are also transformed according to \onee, the $SL(2;R)$ invariance of the formula \bogi\ is easily verified. The above procedure has the advantage that it not only yields the solutions admitting Killing spinors, for given boundary conditions, but also the Killing spinors. For the solutions \dyons\ one finds that $$ \eqalign{ \epsilon &= {1\over\sqrt{2}} H^{-{1\over8}}(H\cos^2\xi + \sin^2\xi)^{-{1\over4}}\Big\{ \big[ (\sin^2\xi + H\cos^2\xi)^{1\over2} + H^{1\over2}\cos\xi \big]^{1\over2} + \cr &\qquad \big[ (\sin^2\xi + H\cos^2\xi)^{1\over2} - H^{1\over2}\cos\xi \big]^{1\over2}\gamma_9\Big\}\epsilon_0\ ,} \eqn\onej $$ where the constant spinor $\epsilon_0$ must satisfy $$ \Gamma_*\gamma_9\epsilon_0 = \epsilon_0\ , \eqn\aonej $$ in order that $\epsilon$ satisfy the constraint \constraint. It follows that the dimension of the space of Killing spinors is half that of the vacuum solution, as anticipated. The expression \onej\ for the Killing spinor $\epsilon$ can be rewritten as $$ \epsilon = e^{{1\over2}\theta\gamma_9}H^{-{1\over8}}\epsilon_0\ , \eqn\conej $$ where $$ \tan\theta = H^{-{1\over2}}\tan\xi\ . \eqn\abonej $$ Note that these spinors vanish at the zeros of $H^{-1}$. In order to show that the $SL(2;R)$ transform of the solutions \dyons, for which $\langle \lambda\rangle \ne i$, are also supersymmetric it suffices to show, as pointed out previously in the context of D=4 dyons [\Ortin], that the conditions \Killing\ are $SL(2;R)$ invariant. Let us denote by $\hat {\cal D}(\lambda, {\cal F})$ the covariant derivative $\hat{\cal D}$ in \oneh, thereby making explicit the dependence of this differential operator on the fields. Under the $SL(2;R)$ transformation of these fields, $\lambda\rightarrow \lambda'$ and ${\cal F}\rightarrow {\cal F}'$ (given explicitly in \onee\ and \onefb), one can show that $$ \hat {\cal D}(\lambda', {\cal F}') = e^{{1\over2}\phi\gamma_9} \hat {\cal D}(\lambda, {\cal F}) e^{-{1\over2}\phi\gamma_9} \eqn\inone $$ where $$ \tan\phi = {-ic(\lambda -\bar\lambda)\over 2d + c(\lambda +\bar\lambda) }\ ; \eqn\intwo $$ i.e. $\hat{\cal D}(\lambda, {\cal F})$ is an $Sl(2;R)$-invariant covariant derivatve. If we take the $Sl(2;R)$ transform of $\epsilon$ to be $$ \epsilon' = e^{{1\over2}\phi\gamma_9}\epsilon\ , \eqn\infour $$ then $$ \hat {\cal D}(\lambda', {\cal F}')\epsilon' = e^{{1\over2}\phi\gamma_9}\hat {\cal D}(\lambda, {\cal F})\epsilon\ , \eqn\ainfour $$ Similarly, if $\chi(\lambda, {\cal F})$ is the spinor of \onei\ then one can show that $$ \chi(\lambda', {\cal F}') = e^{-{1\over2}\phi\gamma_9}\chi(\lambda, {\cal F})\ {}. \eqn\inthree $$ It follows that given background fields and a Killing spinor $\epsilon$ satisfying the conditions \Killing\ for $\langle\lambda\rangle =i$, then the spinor $$ \epsilon' = e^{{1\over2}(\theta+\phi)\gamma_9}H^{-{1\over8}}\epsilon_0 \eqn\ainthree $$ satisfies the same conditions for the $Sl(2;R)$ transformed solution with new asymptotic value $\langle\lambda' \rangle \ne i$. Incidentally, this result establishes the $Sl(2;R)$ invariance of the modified Nester tensor $\hat E^{\mu\nu}$ (assuming the above transformation property of $\epsilon$) and the invariance of the Bogomolnyi bound is an immediate consequence of this. \chapter{Singularity structure} We now turn to the singularity structure of the dyonic membrane solutions \dyons. Near a zero of $H^{-1}$ we have $$ H \sim {\mu \over r^3} \eqn\singone $$ where $$ r\equiv |{\bf x}-{\bf x}_n|\ . \eqn\singtwo $$ The asymptotic metric is $$ r^{3\over2}(-dt^2+dy^2+dz^2) + {dr^2\over r^{3\over2}} + r^{1\over2}d\Omega_4^2 \eqn\singthree $$ where $d\Omega_4^2$ is the metric on the unit 4-sphere. One sees from this result that the proper distance to $r=0$ on a surface of constant $t,y,z$ is finite, and that the radius of the four-sphere of constant $r$ on this surface shrinks to zero as $r\rightarrow 0$. It follows that the `lines' of force of $F$ must end on a singularity at $r=0$. It is instructive to consider the membrane spacetime in the metric $$ d\tilde s^2 = e^{2\sigma}ds^2\ , \eqn\abonej $$ for which $$ d\tilde s^2 = (\cos^2\xi + H^{-1}\sin^2\xi)[-dt^2 + dy^2 + dz^2] + (\sin^2\xi + H\cos^2\xi)\; d{\bf x}\cdot d{\bf x}\ . \eqn\athree $$ The purely electric case now has a timelike naked singularity at zeros of $H^{-1}$, i.e. at a membrane core, so it would have to be identified with a fundamental membrane. For this reason, one might choose to call the metric $d\tilde s^2$ the `membrane metric'. Note that it would be the `string metric' if $\sigma$ were the dilaton, but $\sigma$ is {\it not} the dilaton. In this `membrane metric' the metric for a membrane carrying magnetic charge approaches the asymptotic metric $$ d\tilde s^2 \sim \cos^2\xi\Big\{ [-dt^2 + dy^2 + dz^2] + H\; d{\bf x}\cdot d{\bf x}\Big\} \eqn\bthree $$ near any of the membrane cores. Since $H\sim {\mu\over r^3}$ in this limit, we now find that the proper distance to $r=0$ is infinite on a hypersurface of constant $t,y,z$. Moreover, this remains true for timelike and null geodesics. Thus, the dyonic multi-membrane solutions are geodesically complete {\it in the `membrane' metric} provided that the magnetic charge is non-zero. Because $\sigma$ is not the dilaton, the interpretation of the above result within (type II) string theory is unclear. Moreover, since the dilaton has been set to zero by the truncation, there is no longer any distiction between the Einstein and string metrics. Thus, the fact that our D=8 dyonic membrane solutions are singular in the Einstein metric implies that the string metric is also singular and this must be considered a difficulty in the context of type II superstring theory. Fortunately, this difficulty has a simple resolution if one considers the dyonic solutions as solutions of D=11 supergravity, which can be viewed as an effective action for the strongly coupled type IIA superstring [\PKTb,\Witten]. Consider the following 11-metric and four-form $$ \eqalign{ ds_{11}^2 &= e^{{2\over3}\sigma} ds_8^2 + e^{-{4\over3}\sigma}\; d{\bf u}\cdot d{\bf u}\cr F_{11} &= F + 6du_1\wedge du_2\wedge du_3\wedge d\rho\ , } \eqn\bfour $$ where ${\bf u}$ are the coordinates of $T^3$ and $F$ is a field strength four-form (F=dA) of the eight-dimensional spacetime. This field configuration solves the equations of D=11 supergravity if the 8-metric, four-form $F$, and scalar fields $\sigma$ and $\rho$ solve the D=8 field equations \aonec. This allows us to lift the D=8 dyonic membrane solutions \dyons\ to D=11. The result is $$ \eqalign{ ds_{11}^2 &= H^{-{2\over3}}\Big[\sin^2\xi +H\cos^2\xi\Big]^{1\over3} (-dt^2 + dy^2 + dz^2) \cr &+ H^{1\over3}\Big[\sin^2\xi +H\cos^2\xi\Big]^{1\over3} d{\bf x}\cdot d{\bf x} + H^{1\over3}\Big[\sin^2\xi +H\cos^2\xi\Big]^{-{2\over3}} d{\bf u}\cdot d{\bf u}\cr F_{11} &= {1\over2}\cos\xi (\star dH) +{1\over2}\sin\xi \; dH^{-1}\wedge dt\wedge dy\wedge dz\cr &\qquad -{3\sin 2\xi \over 2[\sin^2\xi + H\cos^2\xi]^2}du_1\wedge du_2 \wedge du_3\wedge dH\ .} \eqn\bfive $$ In the purely electric case, $\cos\xi=0$, we have $$ \eqalign{ ds_{11}^2 &= H^{-{2\over3}}(-dt^2 + dy^2 + dz^2) + H^{1\over3}\big(d{\bf x}\cdot d{\bf x} + d{\bf u}\cdot d{\bf u}\big)\cr F_{11} &= {1\over2}dH^{-1}\wedge dt\wedge dy\wedge dz\ . } \eqn\bsix $$ The harmonic function $H({\bf x})$ can now be interpreted as a harmonic function on $\bb{E}^5\times T^3$. The only difference between this solution of D=11 supergravity and the multi-membrane solution found in [\DS] is that there $H$ was a harmonic function on $\bb{E}^8$. Thus, the solution \bsix\ can be interpreted as a D=11 membrane in a background spacetime of topology $M_6\times T^3$ instead of $M_{11}$, where $M_k$ indicates a $k$-dimensional Minkowski spacetime. In the purely magnetic case, $\sin\xi=0$, we have $$ \eqalign{ ds_{11}^2 &= H^{-{1\over3}}(-dt^2 + dy^2 + dz^2 + d{\bf u}\cdot d{\bf u} ) + H^{2\over3}d{\bf x}\cdot d{\bf x}\cr F_{11}&= {1\over2}\star dH\ , } \eqn\bseven $$ which is the fivebrane solution of D=11 supergravity [\Gu], except for the periodic identification of the $T^3$ coordinates. We can therefore interpret the purely magnetic D=8 membrane as a D=11 fivebrane wrapped around a three-torus. The D=11 multi-fivebrane solution of [\Gu] is geodesically complete [\GHT], the singularities of $H$ being degenerate Killing horizons, so the singularity of the magnetic D=8 membrane solution is resolved by its interpretation in D=11, apart from mild singularities introduced by the periodic identification of the $T^3$ coordinates. These results for the purely electric and purely magnetic D=8 membranes confirm the assumption made in [\PKTc] concerning their D=11 origin. Now we find that the more general dyonic membrane solution also has a D=11 interpretation. Although the D=11 solution does not have an obvious $p$-brane interpretation, it is non-singular, as we now show. Provided the magnetic charge is non-zero, i.e. $\cos\xi\ne0$, the asymptotic form of the metric $ds_{11}^2$ of \bfive\ near any zero of $H^{-1}$ is $$ ds_{11}^2 \sim (\cos\xi)^{2\over3}\Bigg\{ H^{-{1\over3}}(-dt^2 +dy^2 +dz^2 + d{\bf v}\cdot d{\bf v} ) + H^{2\over3}d{\bf x}\cdot d{\bf x}\Bigg\}\ , \eqn\beight $$ where we have set ${\bf u}=(\cos\xi){\bf v}$. Apart from the overall factor the result is independent of $\xi$. That is, the structure of the dyonic membrane near the singularities of $H$ is the same as for the purely magnetic case. We conclude that the singularities of the dyonic membranes are equally resolved in D=11. \chapter{Comments} In this paper we have obtained a bound on the tension of membrane solutions of N=2 D=8 supergravity, and we have found the supersymmetric membrane solutions that saturate this bound. In general these solutions are dyonic. Since N=2 D=8 supergravity is obtained by a $T^3$ compactification of $D=11$ supergravity, followed by a consistent truncation of the massive modes, the D=8 dyonic membranes can be interpreted as solutions of D=11 supergravity. The purely electric and purely magnetic D=8 membranes become the D=11 membrane and fivebrane respectively. The dyonic membranes have no obvious $p$-brane interpretation but they are new solutions of D=11 supergravity which are non-singular if the periodic identification of the $T^3$ coordinates (${\bf u}$) is relaxed. These new solutions are intermediate between the D=11 membrane and fivebrane solutions. They might therefore be expected to play a role in the conjectured D=11 membrane/fivebrane duality [\HT,\PKTc]. The dyonic membrane solutions were given initially for a particular choice of the asymptotic values of the scalar fields that parameterise the possible vacua, but they can then be found for any choice of vacuum by means of an $Sl(2;R)$ transformation. In the context of type II string theory, an infinite set of dyonic membrane solutions can be found, in equivalent vacua, by the action of an $Sl(2;Z)$ subgroup of $Sl(2;R)$ since this is a subgroup of the $SO(2,2;Z)$ T-duality group. As explained in the introduction, this group can be re-interpreted as the S-duality group of the equivalent heterotic string theory after a compactification of the D=8 type II superstring to D=4 on $K_3$. Some D=4 dyon solutions of the heterotic string will thereby acquire an interpretation as D=8 dyonic membranes wrapped around the homology two-cycles of $K_3$. These dyons {\it all} correspond to non-perturbative R-R states in the type II D=4 superstring but, according to the type II/ heterotic equivalence conjecture, correspond to perturbative states of the heterotic string and their non-perturbative S-duals. In fact, they must include the dyons that can become massless at special points in the $K_3$ moduli space [\HTb], as expected from the known symmetry restoration of the heterotic string at special points in its moduli space. Dyonic membranes have many features in common with dyons. For example, let us suppose that there is a purely magnetic membrane with charges $(p,q)=(1,0)$ when $\langle\lambda\rangle=i$; this amounts to the assumption that, in this vacuum, the choice of $\xi=0$ in \dyons\ is admissable in the quantum theory. Now consider a new vacuum related to the original one by an $Sl(2;R)$ transformation with the element $$ \pmatrix{a&b\cr c&d} = \pmatrix{1&b\cr0&1}\ . \eqn\dyonsfive $$ One finds that $\langle\lambda'\rangle = b+ i$, or equivalently $\langle\sigma\rangle =0$, $2\langle\rho\rangle=b$, in the new vacuum and that the membrane solution in this vacuum has charges $(p,q)= (1,b)= (1,2\langle\rho\rangle)$. Thus, a dyonic membrane with unit magnetic charge has a fractional electric charge given by $$ q= 2\langle\rho\rangle \eqn\newone $$ This is just the generalization to dyonic membranes of the Witten effect for dyons [\Witb]. The identification of vacua related by an $Sl(2;Z)$ tansformation implies, in particular, that $2\rho\cong 2\rho +1$, so the value of $q$ for a dyon with unit magnetic charge will change by one as the asymptotic value of $2\rho$ is smoothly continued from $2\langle\rho\rangle$ to $2\langle\rho\rangle +1$. In the D=4 dyon case, this continuation of $\langle\rho\rangle$ can be realized physically by transport around an axion string. In the D=8 dyonic membrane case it could be achieved by transport around an axionic fivebrane. There is, however, a new feature of dyonic membranes not shared by dyons. To see this, we note that given the existence of a particle with charges $(0,1)$ in the vacuum with $\lambda=i$, the DSZ quantization condition implies that for any other particle with charges $(p,q)$, necessarily $p\in Z$, i.e. while electric charge can be fractional, magnetic charge cannot be. Had we assumed the existence of a particle with charges $(1,0)$ we would have instead deduced that $q\in Z$ and $p$ could be fractional. The DSZ quantization condition does not distinguish between these possibilities, but perturbation theory does: in string perturbation theory there exist particles with only electric charge and all semi-classical dyons have integer magnetic charge. A similar conclusion can be made for any of the vacua in the same equivalence class of $\lambda=i$; as we saw earlier for dyonic membranes, the assumption that there exist purely electric solutions is equivalent to the assumption that $\tan\xi$ is rational. It seems, therefore, that for dyons the appeal to perturbation theory allows us to restrict the allowed values of the angular parameter analogous to $\xi$, but the same does not apply to dyonic membranes, at least in the context of type II superstring theory, because all membrane solutions, electric, magnetic or dyonic, are non-perturbative. \vskip 1cm \noindent{\bf Acknowledgements:} G.P. was supported by a Royal Society University Research Fellowship. J.M.I thanks the Commission of the European Community and CICYT (Spain) for financial support. We thank C.M. Hull for helpful discussions. \refout \bye
\section{1. Introduction} The galactic center region has been extensively observed in the molecular lines, particularly in the CO line emission (Oort 1977; Scoville et al 1974; Liszt 1988; Liszt and Burton 1978, 1980; Burton and Liszt 1983, 1993; Brown and Liszt 1984; Heiligman 1987; Bally et al 1987; 1988; Genzel and Townes 1987; Stark et al 1989; G{\"u}sten 1989). Besides the 4-kpc molecular ring, the CO emission is strongly concentrated in the central a few degree (Dame et al 1987). Moreover, the molecular gas in the central region has a strong concentration within $|l|<1^\circ$ (150 pc) where the majority of the nuclear disk gas is confined (Scoville et al 1974; Bally et al 1987; Heiligman 1987). This high concentration of dense interstellar matter in a small region is also clearly visible in the far IR emission (e.g., Cox and Laureijs 1989) and in the CII emission (Okuda et al 1989). The radio continuum emission also indicates a highly concentrated nuclear disk of ionized gas (Altenhoff et al 1978; Handa et al 1987). On the other hand, the region between galactocentric distances $\sim 200$ pc ($l\sim 1\Deg4$) and $\sim 2$ kpc (15$^\circ$) appears almost empty in the CO emission (Bally et al 1987; Knapp et all 1985). The total molecular mass in the $|l|<1^\circ$ region estimated from the CO emission amounts to $ \sim 1.4 \times 10^8 M_{\odot \hskip-5.2pt \bullet}$ for a traditional CO-to-H$_2$\ conversion factor, or, more probably, $\sim 4.6 \times 10^7 M_{\odot \hskip-5.2pt \bullet}$ for a new conversion factor (see section 3). On the other hand, the HI mass within the 1.2 kpc tilted disk ($l<8^\circ$) is only of several $10^6 M_{\odot \hskip-5.2pt \bullet}$ (Liszt and Burton 1980). Hence, we may consider that the central $\sim 1$ kpc region is dominated by the molecular disk of $\sim 150$ pc ($\sim 1^\circ$) radius, outside of which the gas density becomes an order of magnitude smaller. Various molecular gas features in the central $\sim100-200$ pc region have been discussed by various authors, such as a disk related to the 1.2 kpc tilted rotating disk (Liszt and Burton 1980; Burton and Liszt 1992), molecular rings and spiral arms of a few hundred pc scale (Scoville et al 1974; Heiligman 1987; Bally et al 1987), and the expanding molecular ring of 200 pc radius (Scoville 1972; Kaifu et al 1972, 1974). On the other hand, Binney et al (1991) have modeled the ``expanding-ring feature" or the ``parallelogram'' on the $(l, V)$\ (longitude-velocity) plot in terms of non-circular kinematics of gas by a closed orbit model in a bar potential. It is known that the gas in this parallelogram shares only a small fraction of the total molecular mass in the galactic center: the majority of the gas composes more rigid-body like features in the $(l, V)$\ plots. The CO gas in the central 100 - 200 pc regions in nearby galaxies have been observed by high-resolution mm-wave interferometry, and their distribution and kinematics have been extensively studied (e.g., Lo et al 1984; Ishiguro et al 1989; Ishizuki et al 1990a,b). The central gas disks of galaxies appear to comprise spiral arms or circum-nuclear rings of a few hundred pc size. Such a gaseous behavior can be reproduced to some extent by theoretical simulations of accretion of gas clouds in a central gravitational potential (e.g., Noguchi 1988; Wada and Habe 1992). In this paper, we revisit the major part of the nuclear molecular disk ($|l|<\sim 1^\circ$) by analyzing the molecular line data in the premise that the nuclear disk may comprise accretion ring or spiral structures similar to those found in external galaxies. In this paper we reanalyze the data cube of the $^{13}{\rm CO}~(J=1-0)$-line emission observed by Bally et al (1987) with the 7-m off-set Cassegrain telescope of the Bell Telephone Laboratory. The distance to the Galactic Center is assumed to be 8.5 kpc throughout this paper. \section{2. Longitude-Velocity $(l, V)$\ Diagrams} \sub{2.1. Data} The angular resolution of the observations with the Bell-Telephone 7-m antenna at $^{13}$CO line was 1$'.7$. The data used here are in a $(l, b, V_{\rm LSR})$ cube in FITS format. The cube covers an area of $-1^\circ.1 \le l \le 0^\circ.92,~-21' \le b \le 17'$, or 300 pc $\times$ 94 pc region for a 8.5 kpc distance. The velocity coverage is $-250 \le V_{\rm LSR} \le 250$ km s$^{-1}$. The cube comprises 127, 39, and 183 channels at 1$'$, 1$'$, and 2.75 km s$^{-1}$\ intervals, respectively, We also use the CS line data in a $(l,b, V)$ cube with dimensions 151, 42, and 163 channels at intervals 2$'$, 1$'$, and 2.75 km s$^{-1}$, which covers an area of $-1^\circ \le l \le 4^\circ$, $-25' \le b \le 16'$, and $-250\le V_{\rm LSR} \le 190$ km s$^{-1}$. The intensity scale of the data is the main-beam antenna temperature approximately equivalent to brightness temperature in Kelvin. The observational details are described in Bally et al (1987, 1988). We made use of the AIPS and IRAF software packages for the reduction. \sub{2.2. Subtraction of the Local and Foreground Components} In order to analyze molecular gas features in the Galactic Center region, we first subtract contaminations by local and foreground molecular clouds at low velocities. Since $^{13}$CO line is optically thin for the foreground clouds, the contaminations appear as emission stripes superposed on the galactic center emission, The subtraction of foreground emission is essential when we derive the mass and kinetic energy of molecular gas features. Such a ``cleaning" also helps much the morphological recognition of features on the $(l, V)$\ and $(b, V)$\ diagrams. Fig. 1a shows an $(l, V)$\ diagram averaged in a latitude range of $-17' \le b \le 12'$. The diagram is strongly affected by ``stripes'' at a low velocities elongated in the direction of longitude with narrow velocity widths, including the 3-kpc expanding arm at $-52$ km s$^{-1}$. In order to eliminate these stripes, we applied the ``pressing method'' as developed for removing scanning effects in raster scan observations (Sofue and Reich 1979). We briefly describe this method below. \centerline{--Fig. 1a, b --} The original $(l, V)$\ map M$_0$ is trimmed by $-70 \le V_{\rm LSR} \le 50$ km s$^{-1}$\ to yield M$_1$ where the local and foreground gas contribution is significant. The trimmed map M$_1$ is smoothed only in the $V$ direction by 5 channels (14 km s$^{-1}$) using a boxcar or Gaussian smoothing task, yielding M$_2$. Smoothed map M$_2$ is subtracted from M$_1$ to yield M$_3 ~(=M_1 - M_2)$. Map M$_3$ is then smoothed only in $l$ direction by 20 channels (20$'$) (boxcar or Gaussian) to yield M$_4$. This M$_4$ map approximates the contribution from the local gas that is dominated by elongated features in the longitudinal direction. We then subtract M$_4$ from M$_1$ to obtain M$_5~(=M_1-M_4)$. This M$_5$ is, thus, a map in which the local gas contribution has been roughly subtracted. M$_5$ is then smoothed in $V$ direction by 5 channels. Then we replace M$_2$ by this smoothed map, and repeat the above procedures twice (or more times) until we obtain the second (or $n$-th) M$_5$. Finally, the $-70 \le V_{\rm LSR} \le 50$ km s$^{-1}$\ part of the original map M$_0$ is replaced by M$_5$ to yield M$_6$. Now, we have a ``pressed'' map M$_6$ in which corrugations due to local gas clouds have been removed out. Fig. 1b shows the thus obtained map M$_6$ for the same $(l, V)$\ diagram as in Fig. 1a. We have applied this algorithm (the pressing method) to all $(l, V)$\ and $(b, V)$\ diagrams in the cube, and created a new $(l, b, V)$ cube, which is almost free from local and foreground contaminations. In the present paper we use this new cube. We also applied the pressing method to remove scanning effects, which had originated during the data acquisition, in every diagram such as intensity maps in the $(l,b)$ space. By comparing the original and the thus `pressed' maps, we estimated the contribution of the local/foreground emission to be 5\% of the total emission, and 9\% of the emission with $|V_{\rm LSR}|<100$ km s$^{-1}$. Thus, without the subtraction, the mass and energetics would be overestimated by about 5 to 9\%. Moreover, if the gas out of the disk component at $|b|>10'$ is concerned, this local contribution would amount to more than 10\%. Hence, the subtraction of the foreground emissions is crucial in a quantitative discussion of the features discussed in this paper. \sub{2.3. ``Arms'' in Longitude-Velocity $(l, V)$\ Diagrams} Fig. 2 shows $(l, V)$\ diagrams near the galactic plane averaged in $4'$ latitude interval after subtraction of the local/foreground components. Various features found in these diagrams have been discussed in Bally et al (1987, 1988). In this paper we highlight continuous features (ridges) traced in the $(l, V)$\ diagrams. The major structures of the ``disk component'' at low latitude ($|b|<\sim 10'$=25 pc) are ``rigid-rotation'' ridges, which we call ``arms''. Fig. 3 illustrate these ridges (arms) which can be identified in the diagrams as coherent structures. In Table 1 we summarize the identified features, and describe below the individual arms. Heiligman (1987) has used these ridges to derive a rigid rotation curve. At higher latitudes ($|b|>\sim 10'$) the so-called expanding ring features at high velocities ($|V_{\rm LSR} >100$ km s$^{-1}$), which will be discussed in a separate paper. \centerline{-- Fig. 2 --} \centerline{-- Fig. 3 --} \centerline{-- Table 1 --} \sub{2.3.1. Arm I} The most prominent $(l, V)$\ arm is found as a long and straight ridge, slightly above the galactic plane at $b\sim 2'$, which runs from $(l, V)$=($0^\circ.9, 80$ km s$^{-1}$) to ($-0^\circ.7, -150$ km s$^{-1}$), and extends to ($-1^\circ.0, -200$ km s$^{-1}$). This arm intersects the line at $l=0^\circ$ at negative velocity $V_{\rm LSR}=-40$ km s$^{-1}$, indicating that the gas is approaching us at $l=0^\circ$. We call this ridge Arm I. A part of this arm can be traced also below the galactic plane at $b=-0.1$$^\circ$, running from $(l, V)$=(0.8$^\circ$, 60 km s$^{-1}$) to ($0^\circ.1, -20$ km s$^{-1}$). Its positive longitude part is connected to the dense molecular complex Sgr B, which is extended both in space and velocity, from $b=-0.25$ to 0.07$^\circ$\ and $V_{\rm LSR}$=20 to 100 km s$^{-1}$. \sub{2.3.2. Arm II} Another prominent arm is seen at negative latitude at $b\sim -6'$, running from $(l, V)$=(0$^\circ$.1, 60 km s$^{-1}$) to $(-0^\circ.6, -80$ km s$^{-1}$). We call this ridge Arm II. It is bent at $l\sim 0^\circ.1$ and appears to continue to $(l, V)$=(1$^\circ$, 100 km s$^{-1}$), and merges with Arm I at the Sgr B complex region. The negative longitude part also merges with Arm I, and is connected to the Sgr C complex. Arm II intersects $l=0^\circ$ at positive velocity of $V_{\rm LSR}=50$ km s$^{-1}$. \sub{2.3.3. Arms III and IV} At positive latitude ($b\sim 0^\circ.01$ to $0^\circ.2$), another arm can be traced running from $(l, V)$=(0$^\circ$, 140 km s$^{-1}$) to ($-0^\circ.15$, 10 km s$^{-1}$). Its counterpart to the negative longitude side appears to be present at $(l, V)$=($-0^\circ.45, -120$ km s$^{-1}$) to ($-0^\circ.55, -180$ km s$^{-1}$). We call this ridge Arm III. Bally et al (1988) called this the ``polar arc'', and discussed its connection to Sgr A. A branch can be traced from $(l, V)$=(0$^\circ$.1, 60 km s$^{-1}$) to ($0^\circ, -20$ km s$^{-1}$), apparently being bifurcated from Arm II at $l\sim 0^\circ.1$. This ridge intersects $l=0^\circ$ at negative velocity ($V_{\rm LSR} -50$ km s$^{-1}$). We call this ridge Arm IV. \sub{2.4. ``Rigid-rotation'' in $(l, V)$\ Plane and ``Arms and Ring'' in the Galactic Plane} We emphasize that ``rigid-rotation'' ridges in $(l, V)$\ diagrams for edge-on galaxies, whose rotation curves are usually flat, are generally interpreted as due to spiral arms and rings. Indeed, the rigid-rotation ridge in the CO $(l, V)$\ diagram of the Milky Way is identified with the 4-kpc molecular ring (e.g., Dame et al 1987; Combes 1992). Many edge-on spiral galaxies like NGC 891 are found to show similar $(l, V)$\ ridges in HI and CO, which are also interpreted to be spiral arms and rings (e.g., Sofue and Nakai 1993, 1994). The circular rotation velocity as defined by $V_{\rm rot}=(R \partial \Phi / \partial R)^{1/2}$, remains greater than at least 150 km s$^{-1}$\ from the nuclear few pc region till the 1 kpc radius region (Genzel and Townes 1987). Here, $\Phi$ is the potential and $R$ is the distance from the nucleus. Hence, the actual rotation should not be rigid at all: The rigid-rotation ridges in the $(l, V)$\ plane such as Arms I to IV in the Galactic Center can thus be more naturally attributed to real arms and rings. \section{3. Intensity Distribution and the Galactic Center Arms} \sub{3.1. Intensity Maps and Masses} Fig. 4a shows the total intensity map integrated over the full range of the velocity ($-250 \le V_{\rm LSR} \le 250$ km s$^{-1}$). This map is about the same as that presented by Stark et al (1989), except that the local gas has been removed. Fig. 4b shows the same in grey scale and a that with the vertical scale in $b$ direction enlarged twice. \centerline{-- Fig. 4 --} First of all the intensity map can be used to obtain the molecular mass. However, the conversion of the CO intensity to H$_2$\ mass is not straightforward. We have recently studied the correlation of the conversion factor $X_{12}$ for the $^{12}$CO$(J=1-0)$ line with the metal abundance in galaxies (Arimoto et al 1994). We have obtained a clear dependency of $X$ on the galacto-centric distance $R$ within individual galaxies, which is almost equivalent to the metallicity dependence. For the Milky Way we have $$ X_{12}(R)=0.92 (\pm 0.2)\times 10^{20}{\rm exp} (R/R_{\rm e})$$ where $R_{\rm e}=7.1$ kpc is the scale radius of the disk. Applying this relation to the Galactic center, we obtain a conversion factor at the Galactic center as $X_{12}=0.92(\pm 0.2)\times10^{20}~{\rm [H_2~cm^{-2}/K~kms~s^{-1}]}$, about one third of the solar vicinity value. We then assume that the $^{12}$CO and $^{13}$CO intensities are proportional, and estimate the ratio by averaging observed intensity ratios for the inner Galaxy at $l \le 20^\circ$ (Solomon et al 1979); $I_{\rm 12CO}/I_{\rm 13CO} \simeq 6.2\pm 1.0$. Then, we obtain a conversion factor for the $^{13}$CO line intensity in the galactic center region as $X_{13}(R=0)\simeq 5.7 \times 10^{20} ~{\rm [H_2~cm^{-2}/K~kms~s^{-1}]}$, and we use this value throughout this paper. The correction factor from the H mass to real gas mass is given by $\mu=1/X=1.61$, where $X$ is the hydrogen abundance in weight. Here, the following relation has been adopted (Shaver et al 1983): $Y=0.28+(\Delta Y/\Delta Z) Z = 0.34$ in weight, where $Z=0.02$ is the abundance of the heavy elements and $\Delta Y/\Delta Z=3$ is the metallicity dependence of the helium abundance $Y$ in the interstellar matter, and so, the hydrogen abundance is $X=0.62$. So, the surface mass density of molecular gas after correction for the mean weight of gas is given by $$ \sigma \sim 14.6 (\pm 3) ~ I/\eta~[ M_{\odot \hskip-5.2pt \bullet} {\rm pc}^{-2}], \eqno(2) $$ where $$ I\equiv\int T_{\rm A}^* dv {\rm~ [ K~ km~s^{-1}}] \eqno(3) $$ is the integrated intensity of $^{13}{\rm CO}(J=1-0)$ line emission and $\eta=0.89$ is the primary beam efficiency of the antenna. The total mass of molecular gas (including He and metals) can be estimated by $$ M~ [M_{\odot \hskip-5.2pt \bullet}]= 14.6 \int I/\eta dx dy~ [{\rm K~km~s^{-1}~pc^2}]. \eqno(4) $$ Using these relations, we have estimated the total molecular gas mass in the observed area ($-1^\circ.0 \le l \le 0^\circ.92,~ -21' \le b \le 17')$ after removing the local and foreground contribution to be $ 4.6 (\pm 0.8) \times 10^7 M_{\odot \hskip-5.2pt \bullet}$. We have also estimated the total molecular mass of the ``disk'' component, which comprises most of the ridge-like features in the $(l, V)$\ diagrams, excluding the expanding ring feature (or the parallelogram) at high velocities ($|V_{\rm LSR}| >\sim 100 - 150$ km s$^{-1}$). The disk component has the mass $3.9\times10^7 M_{\odot \hskip-5.2pt \bullet}$, which is 85\% of the total in the observed region. On the other hand, the expanding ring (or the parallelogram) shares only $6.7\times 10^6 M_{\odot \hskip-5.2pt \bullet}$ (15\%) in the region at $|l|<1^\circ$. \sub{3.2. Ring and Arms in Intensity Maps} In order to clarify if each of the arms traced in the $(l, V)$\ diagrams (Fig. 1-3), particlurly Arms I and II, is a single physical structure in space, we have obtained velocity-integrated intensity map in the $(l, b)$\ plane for each of the arms. Thereby, we integrated the CO intensity in the velocity ranges as shown in Fig. 5 individually for Arms I and II. Fig. 6 show the integrated intensity maps corresponding to Arms I and II, together with a summation of I and II. In Table 1 we summarize the derived parameters. \centerline{-- Fig. 5 --} \centerline{-- Fig. 6 --} \sub{3.2.1. Galactic Center Arms I, II} In the intensity map, Arm I can be traced as a single, thin arc-like arm from $l=0^\circ.9$ near the Sgr B complex toward negative longitude at $l=-1^\circ.0$. We call this spatial arm the Galactic Center Arm I (GCA I). The angular extent is as long as $1^\circ.9$ (280 pc) in the longitudinal direction, whereas the thickness in the $b$ direction is as thin as $\sim 5'$ (13 pc; see section 3.2.2). The Sgr B molecular complex is much extended in the $b$ direction by about $0^\circ.4$ (60 pc), and composes a massive part of the arm. A ``return'' of this arm can be traced from $(l, b)$=$(0^\circ.9, 0^\circ)$ to $(0^\circ.2, -0^\circ.07)$, and is more clearly recognized in Fig. 4 at $V=83 \sim 167$km s$^{-1}$. This can be also clearly seen in the $(l, V)$\ diagram in Fig. 2 at $b\sim -0^\circ.1$. In the negative $l$ side, the arm appears to be bifurcated at $l\sim -0^\circ.65$, and linked to Arm II. This can be more clearly observed in Fig. 4 at $V=-83 \sim 0$ km s$^{-1}$. The intensity in Fig. 6a has been integrated to give a total mass of molecular gas involved in GCA I (in the velocity range as shown in Fig. 5a) to be $M = 1.72 \times 10^7 M_{\odot \hskip-5.2pt \bullet} $. Arm II can be traced as a a single bright ridge from $l\sim 0^\circ.3$ to $-0^\circ.7$, and the thickness is about 6$'$ (15 pc), and makes GCA II. The mass of Arm II is estimated to be $ 1.35 \times 10^7 M_{\odot \hskip-5.2pt \bullet}$. Thus, the total mass involved in GCA I and II is estimated to be $ 3.07 \times 10^7 M_{\odot \hskip-5.2pt \bullet}$, and shares almost 67\% of the total gas mass in the observed region, and 78\% of the disk component. \sub{3.2.2. The 120-pc Molecular Ring} As shown in Fig. 4 and 6, GCA I and II compose a global ring structure, which is tilted and slightly bent. If we fit the GCA I and II by a ring, its angular extent in the major axis is $1^\circ.6$ from $l=-0^\circ.7$ to $0^\circ.9$, and so, the major axis length (diameter) is 240 pc, and the radius 120 pc. The minor axis length is estimated to be $7'.9$ from the maximum separation between Arm I and II at $l\sim -0^\circ.2$ (see Fig. 7). Therefore, the inclination of the I+II ring is $i=85^\circ$ from the minor-to-major axis ratio. The center of the ring, as fitted by the above figures, is at $(l, b)$=$(0^\circ.1,0^\circ.0)$ We call this ring the 120-pc Molecular Ring. {}From these we conclude that the spatial distribution of the molecular gas associated with the principal ridges in the $(l, V)$\ diagrams comprises a circum-nuclear ring of radius $R\simeq 120$ pc inclined by 5$^\circ$\ from the line of sight. \sub{3.2.3. Cross Section of the Arms} Fig. 7 shows the intensity variation perpendicular to the galactic plane across GCA I and II averaged from $l=0^\circ.24$ to $-0^\circ.33$, where the arms are most clearly separated. Since the effective resolution of the present data is $(\theta^2+\Delta b^2)^{1/2}=2'.0$, where $\theta=1'.7$ is the beam width and $\Delta b=1'.0$ is the grid interval, the arms are sufficiently resolved. The two peaks in the figure at $b=1'.8$ (Arm I) and $b=-6'.0$ (Arm II) can be fitted by a Gaussian intensity distributions as $(T_{\rm B,~ peak}, {\rm FWHM}) = (0.27{\rm ~K}, 5'.3)$ and $(0.33{\rm ~K}, 5'.5)$, respectively. Namely, the arms are as thin as 13.0 (GCA I) and 13.5 pc (GCA II). If we subtract the contributions from these two arm components, the residual intensity in the whole area in Fig. 7 is only 36\% of the total intensity. The intensity coming from the inter-arm region between the arms shares only 12\% of the intensity from the two arms. This would be an upper limit, as the region displayed in this figure is the weakest part along the arms without any significant molecular clumps and condensations. Thus, we conclude that the molecular gas as observed in the CO line emission in the region discussed in this paper is almost totally confined within the two major arms. Therefore, the central 100 pc radius region is almost empty, making a hole of molecular gas, except the nuclear few pc region surrounding Sgr A. \centerline{-- Fig. 7 --} \sub{3.3. Velocity Field} Fig. 8a shows a velocity field as obtained by taking the first moment of the $(V_{\rm LSR}, l, b)$ cube, and therefore, an intensity-weighted velocity field. A general rotation characteristics is clearly seen along the major axis of the ring feature at $b\simeq -6'$. Sgr C molecular spur is seen as a negative velocity spur extending toward negative $b$. GCA III is seen as the tilted high-velocity plume at $(l,b) \sim (0^\circ.2, 0^\circ.1)$. In addition to these individual velocity structures, a large-scale velocity gradient in the latitude direction is prominent in the sense that the positive $b$ side has positive velocity and negative $b$ side negative velocity. This can be attributed to the fact that the high-velocity expanding shell (ring) is more clearly seen in positive velocity at $b>0^\circ$, while the negative velocity part more clearly at $b<0^\circ$ (see section 4). This can be explained by a tilted nature of the expanding oblate molecular shell, as will be discussed in section 4 based on an analysis of $(b, V)$\ diagrams. In fact, if we construct a velocity field, excluding the expanding ring features, we obtain a rather regular velocity field as shown in Fig. 8b. \centerline{-- Fig. 8 --} \sub{3.4. Possible Models for the Galactic Center Arms and Ring} We here try to reproduce the $(l, V)$\ diagram based on a simple spiral arm model. According to the galactic shock wave theory (Fujimoto 1966; Roberts 1969) and the bar-induced shock wave theory (Sorensen et al 1976; Huntley et al 1978; Roberts et al 1979; Noguchi 1988; Wada and Habe 1992), flow vectors of gas in the densest part along the shocked arms are almost parallel to the potential valley that is rigidly rotating at a pattern speed slower than the galactic rotation. In such shocked flows, the gas cannot be on a closed orbit, but is rapidly accreted toward the center along deformed spirals. As the simplest approach to simulate an $(l, V)$\ diagram, we assume that the flow vector of gas is aligned along a spiral with a constant velocity equal to the rotation velocity in the potential. Fig. 9a shows a model, where we have assumed two symmetrical spiral arms with a pitch angle $p=10^\circ$. In addition to a constant circular rotation of gas ($V_{\rm rot}=$constant; flat rotation curve), radial infall motion of $V_{\rm rot} {\rm sin}~p$ is superposed, so that the gas is flowing along the arms into the central region. The density distribution in the arms are shown by the spiral-like contours. The azimuthally averaged density of gas has a hole at the center, or it corresponds to a ring distribution of gas on which two arms are superposed. A calculated $(l, V)$\ diagram is shown by the superposed contours with a tilted X shape. The characteristic features in the observed $(l, V)$\ diagrams can be now qualitatively reproduced. Fig. 9b and 10c show cases where the spiral arms are oval in shape whose major axis is inclined by $\pm 30^\circ$ from the nodal line. Such a case may be expected when the oval potential or a bar in the center is deep enough to produce a non-circular motion. Fig. 9d-f are the same, but the density distribution along the arms has the maximum at the center and the pitch angle is taken larger: $p=20^\circ$. Again, the case of a circular rotation appears to reproduce the observation, while the oval orbit cases result in more complicated $(l, V)$\ plots than the observation. Among these models, the case shown in Fig. 9a or 9b appears to reproduce the observed characteristics in the $(l, V)$\ plot (e.g. Fig. 3) reasonably well. The model in Fig. 9d or 9e with the averaged gas density increasing toward the center may explain observed Arms III and IV. However, the cases corresponding to Fig. 9c and 9f may be excluded. \centerline{-- Fig. 9 --} According to the galactic shock wave theory in a spiral density wave or a bar potential, the shocked gas looses its azimuthal velocity so that the $(l, V)$\ behavior becomes closer to the potential's pattern speed. As the consequence, the apparent rotation velocity of the gas along shocked spiral arms is smaller than that from the rotation velocity. This may be the reason why the observed maximum velocities of the rings/arms (e.g., Arm I near Sgr B) are less than that expected from the gravitational potential. \sub{3.5. Deconvolution into Projection on the Galactic Plane: A Face-on View} We may thus assume that the molecular gas is on a ring or spiral arms whose pitch angle is not so large. Then, it is possible to deconvolve the $(l, V)$\ diagram into a spatial distribution in the galactic plane by assuming an approximately circular rotation. Thereby, we make use of the velocity-to-space transformation (VST), which has been extensively applied to derive the HI gas distribution in our Galaxy (Oort et al 1957). Suppose that a gas element is located at a projected distance $x~ (\simeq l\times 8.5$ kpc) along the galactic plane from the center of rotation, and has a radial velocity $v$. If the rotation is circular at velocity $V_0$, the line of sight distance $y$ of the element from the nodal line can be calculated by $$ y = \pm |x| \sqrt{ \left( V_0 \over v \right)^2 -1} . \eqno(5)$$ The signs must be opposite for Arms I and II. Here, we assume that Arm I is near side, and Arm II far side, so that the signs are $-/+$, respectively. The center of rotation is assumed to be at Sgr A, and $v$ is measured from the intersection velocity at $l=-0^\circ.06$ on each arm ridge. Fig. 10 shows a thus obtained ``face-on'' map of the molecular gas for $V_0=150$ km s$^{-1}$. The arms appear to construct a circum-nuclear ring of radius $\sim 120$ pc. Here, we used $(l, V)$\ diagrams averaged within latitude ranges $-2' \le b \le 6'$ for Arm I and $-5' \le b \le 3'$ for Arm II, so that vertically extended clumps such as Sgr B complex are only partly mapped in this figure. During the deconvolution, we used only the arm component concentrated near the ridges within $\pm 20$ km s$^{-1}$\ in velocity (as illustrated in Fig. 5). Diffuse gas and clumps with velocities far from the arms are not taken into account. The same VST was applied to the HII regions Sgr B1, B2 and C using their H recombination line velocities (Downes et al 1980). We plotted their positions in Fig. 10. The HII regions lie along the arms associated with the molecular complexes, though slightly avoiding the molecular gas peaks. Sgr B and C appear to be at symmetrically opposite locations with respect to the nucleus. We have assumed that Arm I is near side. However, in this kind of simple deconvolution, we cannot distinguish the exact orientation, as is the case of deconvolution of gas distribution inside the solar circle from kinematical information. Hence, it may be possible to assume an opposite configuration of the arm locations: Arm I in far side, and Arm II in near side. \centerline{-- Fig. 10 --} The connection of Arms I and II is not clear from this deconvolution. This is mainly because of the ambiguous position determination near the node, which arises from unknown precise rotation curve. The error is also large at $|l| < \sim 0^\circ.1$, where we applied interpolation from both sides along each arm. Obviously, this kind of deconvolution is not unique, but it was possible here because of the separation of Arms I and II in the $(l,b)$ plane. Therefore, this deconvolution should be taken as a possible hint to the spatial distribution of gas. \sub{3.6. Comparison with Other Galaxies and Models} Accretion spirals, either shocked or not, and rings of molecular gas have been indeed observed in the CO line in many extragalactic systems such as IC 342 (Lo et al 1984; Ishizuki et al 1990a) and NGC 6946 (Ishizuki et al 1990b). The ring structure of molecular gas of 100 to a few hundred pc size is commonly observed in the central regions of spiral galaxies (Nakai et al 1987; Ishiguro et al 1989). See Sofue (1991) for a more number of galaxies with a nuclear molecular ring. Thus, the ring/spiral structure of molecular gas of radius 120 pc in the Milky Way, would be similar to the situation found in external galaxies. There have been various numerical simulations of the accretion of gas toward the central region in spiral and oval potential by gas-dynamical simulations (Sorensen et al 1976; Huntley et al 1978; Roberts et al 1979; Noguchi 1988; Wada and Habe 1992). The models predict a rapid accretion of gas along spiral orbits, and the gas behavior in these models somehow mimic the models illustrated in Fig. 9. A number of simulations of position-velocity diagrams along the galactic plane have been constructed and compared with the observations, in order to understand larger-scale $(l, V)$\ diagrams for our Galaxy both in HI and CO (Mulder and Liem 1986; Liszt and Burton 1978; Burton 1988). Position-velocity diagrams for extragalactic edge-on galaxies in CO have been extensively studied (Sofue and Nakai 1993, 1994; Sofue 1994) and a numerical simulation has been attempted to reproduce the PV characteristics based on the gas dynamics in an oval potential (e.g., Mulder and Liem 1986; Wada et al 1994). Binney et al (1991) have noticed the ``parallelogram'' and calculated theoretical $(l, V)$\ diagrams, and have shown the presence of a bar of 2 kpc length in the Galactic bulge. However, the parallelogram (the expanding ring feature) in Fig. 1b shares only 15\% of the total emission. However, we emphasize that the major structures, which contain 85\% of the molecular mass within 150 pc of the center, are due to the Arms discussed above. \sub{3.7. Relationship with Radio Sources} Fig. 11 shows superposition of a 10-GHz radio continuum map (Handa et al 1987) on the $^{13}$CO and CS intensity maps. We here briefly comment on a global relationship of the major radio sources with molecular features at a spatial resolution of a few arc minutes. Detailed internal structures of individual sources are out of the scope of the present paper, for which the readers may refer to a review by Liszt (1988). \sub{3.7.1. Sgr A} The relationship of molecular features of scales less than a few arc minutes with Sgr A has been discussed by many authors (e.g., Oort 1977; Bally et al 1987; G{\"u}sten 1989). However, these nuclear features, which are of $1'~(\sim 3$ pc) scales, are not well visible in the present plots in so far as the $(l, V)$\ plots are concerned. We only mention that Arm III is a largely tilted out-of-plane plume with high positive velocity, which Bally et al (1988) called the polar arc. Arm IV shows also large velocity gradient, and appears to be an object related to a deep gravitational potential around the nucleus. \sub{3.7.2. Sgr B} The molecular complex at $l\sim 0^\circ.6-0^\circ.9$ on Arm I is associated with the star forming regions Sgr B1 at $(l,b)=(0^\circ.519,-0^\circ.050)$, and Sgr B2 at $(0^\circ.670, -0^\circ.036)$. Sgr B1 and B2, whose radial velocities in H recombination line emission are $V_{\rm LSR}$=45 and 65 km s$^{-1}$, respectively (Downes et al 1980), are also located in the $(l, V)$\ plane at the upper (higher-velocity) edges of molecular clumps. Thus, the de-convolved positions of these continuum sources are slightly displaced from the de-convolved arm, as indicated in Fig. 10. The molecular gas distribution is highly extended in the direction of latitude for about $0^\circ.4$ (60 pc), largely shifted toward the lower side of the galactic plane ($b <0^\circ$). This complex is also much extended in the velocity space: the velocity dispersion amounts to as high as 50 km s$^{-1}$. The internal structure of Sgr B molecular complex has been discussed in detail in relation to the star formation activity, and it was shown that the molecular gas is distributed in a shell, spatially surrounding the continuum peak (Bally et al 1988; Sofue 1990; Hasegawa et al 1993). The present ring model is consistent with the CII line $(l, V)$\ diagram as obtained by Okuda et al (1989), which indicates a rotating ionized gas feature with Sgr B and C on the tangential points of the ring. \centerline{-- Fig. 11 --} \sub{3.7.3. Sgr C} The star forming region Sgr C is associated with a molecular complex, and is located on Arm II at $l\sim -0^\circ.6$. However, the spatial proximity is less significant than that for Sgr B: The radio continuum peak of Sgr C, $(l,b,V_{\rm LSR})=(-0^\circ.57,-0^\circ.09)$, is located at the western edge of the molecular complex, but displaced by about 6$'$(15 pc) from the molecular peak. The LSR velocity of the H recombination line also agrees with the molecular gas velocity, and so, it is located on the de-convolved arm in Fig. 10. The molecular gas in this complex is extended vertically, and molecular spurs are found to extend both toward positive and negative latitude directions. We emphasize that the positive-latitude spur is clearly associated with the inner edge of the western ridge of the Galactic Center Lobe observed in the radio continuum emission (Sofue and Handa 1984; Sofue 1985), as is shown in Fig. 11. \sub{3.7.4. Orbital Displacement vs Alignment of Star Forming Regions and Molecular Arms} The close association of Sgr B and C with GCA I and II may have a crucial implication for the orbits of gas and stars: If the arms are shock lanes in a bar during a highly non-circular motion, the HII regions of a million years old should already be displaced from the molecular arms. Therefore, the fact that Sgr B and C are still near the gas complexes from which they may have been born (after one or more rotations) can be explained only if the stars and gas are circularly co-rotating in the arms at a small pitch angle. This would argue for the validity of the deconvolution process applied in section 3.5. Consider a spiral arm which is a shocked density wave. Star formation from a molecular cloud will be triggered in the arms. It will take about $t \sim 10^6$ years for proto stars to form and shine as OB stars, and therefore, until HII regions are produced. On the other hand, the rotation period of the stars is only $\sim 10^6$ years for $r=100$ pc and $V_{\rm rot}=200 $ km s$^{-1}$. According to the density wave theory, the velocity difference between the rotation velocity and the shocked gaseous arm, which is about the same as the pattern speed of density wave, is of the order of $$V_{\rm rot}-V_{\rm p}=(\Omega_{\rm rot}-\Omega_{\rm p})r. \eqno(6)$$ The azimuthal phase difference between the HII region and the gaseous arm is then $$\Delta \phi \sim (\Omega_{\rm rot}-\Omega_{\rm p})t . \eqno(7) $$ The phase difference for Sgr B2 and its corresponding molecular peak in Fig. 10 (darkest part in Arm I) is roughly $\Delta \phi \sim 5^\circ$, and a similar value is found for Sgr C. If $t\sim 10^6$ yr, we obtain $\Omega_{\rm rot} - \Omega_{\rm p}\sim 0.1$ radian/$10^6$ years $\sim 100$ km s$^{-1}$~kpc$^{-1}$. This is an order of magnitude greater than the value near the solar circle ($\sim 10$ km s$^{-1}$~kpc$^{-1}$). For older HII regions (weaker radio sources) the phase difference would be much greater. Moreover, orbits of stars, and therefore, HII regions, are no longer closed, and must be largely displaced from the orbits of gas. Thus, the HII regions in the central 100 pc of the Galaxy, except for young cases as Sgr B2, would not be associated with molecular gas arms. This will simply explain why the molecular gas features are not directly correlated with the weaker radio sources in the Galactic center (Fig. 11). \section{4. Discussion} By analyzing the $^{13}$CO line BTL data cube, we have shown that most (85\%) of the total molecular gas within $|l|<1^\circ$ comprises rigid-body-like structures in the $(l, V)$\ diagrams, which can be attributed to arms on a ring. Moreover, 66\% of the total gas in the region, and 78\% of the disk component ($|b|<\sim 10'$=25 pc), was found to be confined in the two major Arms I and II. The spiral/ring structures are consistent with the picture drawn by Scoville et al (1974) based on the earlier data, while the scale obtained here is slightly smaller. The structures will be common in external galaxy nuclei in the sense that the gas distribution is spiral- and ring-like. Numerical simulations for a few kpc scale disks have suggested that the features would be understood as the consequence of spiral accretion by a density wave in an oval potential, either shocked or not. Based on qualitative consideration, we have suggested possible models to explain the observed $(l, V)$\ features as shown in Fig. 9a. The molecular mass in the Galactic Center has been derived by usin the most recent CO-to-H$_2$ conversion factor about one third of the conventional value, which has been obtained by detailed analyses of the dependency on the metallicity as well as on the galacto-centric distance (Arimoto et al 1994). This has resulted in a factor of three smaller mass and energetics than the so far quoted values in the literature: The molecular mass within 150 pc radius from the center is estimated to be only $3.9\times 10^7 M_{\odot \hskip-5.2pt \bullet}$. Thus, the molecular gas mass is only a few percent of the total mass in the region estimated as $M_{\rm dyn}=R V_{\rm rot}^2/G \sim 8 \times 10^8 M_{\odot \hskip-5.2pt \bullet}$ for a radius $R \sim 150$ pc and rotation velocity $V_{\rm rot} \sim 150$ km s$^{-1}$. This implies that the self-gravity of gas is not essential in the galactic center, and a given-potential simulation would be sufficient to theoretically understand the region. The expanding molecular ring (or the parallelogram) was shown to share only 15 percent of the total gas mass within the central 1$^\circ$\ region. This feature has been shown to be extending vertically over $\sim 100$ pc above and below the galactic plane (Sofue 1989). For the very different $b$ distribution, it is a clearly distinguished structure from the arms and the ring described in this paper. On the $(l, V)$\ plot, the feature can be fitted by an ellipse of radius 1$^\circ$.2 (Bally et al 1987), slightly larger than the disk discussed in this paper . There have been controversial interpretations about this feature: either it is due to some explosive event (Scoville et al 1972; Kaifu et al 1972, 1974) or due to non-circular rotation of disk gas (Burton and Liszt 1992; Binney 1991). We will discuss this feature based on the present data in a separate paper. \v\v {\bf Acknowledgement}: The author would like to express his sincere thanks to Dr. John Bally for making him available with the molecular line data in a machine-readable format. \section{References} \parskip=0pt \def\hangindent=1pc \hangafter=1 \no{\hangindent=1pc \hangafter=1 \noindent} \hangindent=1pc \hangafter=1 \no Altenhoff, W. J., Downes, D., Pauls., T., Schraml, J. 1979, AAS {35}, 23. \hangindent=1pc \hangafter=1 \no Arimoto, N., Sofue, Y., Tsujimoto, T. 1994, in preparation. \hangindent=1pc \hangafter=1 \no{Bally, J., Stark, A.A., Wilson, R.W., and Henkel, C. 1987, ApJ Suppl 65, 13.} \hangindent=1pc \hangafter=1 \no{Bally, J., Stark, A.A., Wilson, R.W., and Henkel, C. 1988, ApJ 324, 223.} \hangindent=1pc \hangafter=1 \no Binney, J.J., Gerhard, O.E., Stark, A.A., Bally, J., Uchida, K.I., 1991 MNRAS 252, 210. \hangindent=1pc \hangafter=1 \no{Brown, R.L, and Liszt, H.S. 1984, ARAA 22, 223.} \hangindent=1pc \hangafter=1 \no Burton, W. B. 1988, in Galactic and Extragalactic Radio Astronomy, ed. G. L. Verschuur and K. I. Kellermann, 2nd edition (Springer-Verlag, New York) p 295. \hangindent=1pc \hangafter=1 \no Burton, W. B., and Liszt, H. S. 1983, in Surveys of the Southern Galaxy, ed. W. B. Burton and F. P. Israel (Reidel Pub. CO, Dordrecht), p. 149. \hangindent=1pc \hangafter=1 \no Burton, W. B., and Liszt, H. S. 1992 AAS 95, 9. \hangindent=1pc \hangafter=1 \no Combes F 1992 ARAA, 29, 195. \hangindent=1pc \hangafter=1 \no Cox, P., Laureijs, R. 1989, in The Center of the Galaxy (IAU Symp. 136), ed. M.Morris (D.Reidel Publ. Co., Dordrecht) p. 121. \hangindent=1pc \hangafter=1 \no{Dame, T. M., Ungerechts, H., Cohen, R. S., de Geus, E. J., Grenier, I. A., et al. 1987 ApJ 32, 706 \hangindent=1pc \hangafter=1 \no Downes, D., Wilson, T. L., Beiging, J., Wink,J. 1980, AA Suppl. 40, 379. \hangindent=1pc \hangafter=1 \no Fujimoto, M. 1966, in Non-stable Phenomena in Galaxies, IAU Symp. No 29, ed. Arakeljan (Academy of Sciences of Armenia, USSR), p.453. \hangindent=1pc \hangafter=1 \no{Genzel, R., and Townes, C.H 1987, ARAA 25, 377.} \hangindent=1pc \hangafter=1 \no G{\"u}sten, R. 1989, in The Center of the Galaxy (IAU Symp. 136), ed. M.Morris (D.Reidel Publ. Co., Dordrecht) p. 89. \hangindent=1pc \hangafter=1 \no{Handa, T., Sofue, Y., Nakai, N. Inoue, M., and Hirabayashi, H. 1987, PASJ 39, 709.} \hangindent=1pc \hangafter=1 \no Hasegawa, T., Sato, F., Whiteoak, J. B., Miyawaki, R. 1993, ApJ 419, L77. \hangindent=1pc \hangafter=1 \no Heiligman, G. M. 1987 ApJ 314, 747. \hangindent=1pc \hangafter=1 \no{Huntley, J. M., Sanders, R. H., and Roberts, W. W., 1978, ApJ 221, 521.} \hangindent=1pc \hangafter=1 \no{Ishiguro, M., Kawabe, R., Morita, K.-I., Okumura, S. K., Chikada, Y. et al. 1989, ApJ 344, 763. } \hangindent=1pc \hangafter=1 \no{Ishizuki, S. Kawabe, R., Ishiguro, M., Okumura, S. K., Morita, K-I., et al. 1990a Nature 344, 224. \hangindent=1pc \hangafter=1 \no{Ishizuki, S., Kawabe, R., Ishiguro, M., Okumura, S. K., Morita, K. -I. et al. 1990b ApJ 355 436. \hangindent=1pc \hangafter=1 \no{Kaifu, N., Iguchi, T., and Kato, T. 1974, PASJ 26, 117.} \hangindent=1pc \hangafter=1 \no{Kaifu, N., Kato, T., and Iguchi, T. 1972, Nature 238, 105.} \hangindent=1pc \hangafter=1 \no Knapp, G. R., Stgark, A. A., Wilson, R. W. 1985 AJ 90, 254. \hangindent=1pc \hangafter=1 \no Liszt, H. S. 1988 in Galactic and Extragalactic Radio Astronomy, ed. G. L. Verschuur and K. I. Kellermann, 2nd edition (Springer-Verlag, New York) p 359. \hangindent=1pc \hangafter=1 \no Liszt,H. S., Burton, W. B. 1978 ApJ 226, 790. \hangindent=1pc \hangafter=1 \no Liszt, H. S., and Burton, W. B. 1980 ApJ 236, 779. \hangindent=1pc \hangafter=1 \no{Lo, K. Y., Berge, G. L., Claussen, M. J., et al. 1984, ApJ 282, L59. \hangindent=1pc \hangafter=1 \no Mulder, W.A., Liem, B.T., 1986, AA 157, 148 \hangindent=1pc \hangafter=1 \no{Nakai, N., Hayashi, M., Handa, T., Sofue, Y., Hasegawa, T., and Sasaki, M., 1987, PASJ 39, 685.} \hangindent=1pc \hangafter=1 \no Okuda, H., Shibai, H., Nakagawa, T., Matsuhara, T., Maihara, T., et al. 1989 in The Center of the Galaxy (IAU Symp. 136), ed. M.Morris (D.Reidel Publ. Co., Dordrecht) p. 145. \hangindent=1pc \hangafter=1 \no Oort, J. H., Kerr, F. J., Westerhout, G. 1958, MNRAS 118, 379. \hangindent=1pc \hangafter=1 \no Oort, J. H. 1977, ARAA 15, 295 AA Suppl 58, 197.} \hangindent=1pc \hangafter=1 \no Roberts, W. W. 1969, ApJ 158, 123. \hangindent=1pc \hangafter=1 \no Roberts, W. W., Huntley, J. M., van Albada, G. D. 1979, ApJ, 233, 67. \hangindent=1pc \hangafter=1 \no Shaver, P. A., McGee, R. X., Newton, L. M., Danks, A. C., Pottasch, S. R. 1983, MNRAS, 204, 53. \hangindent=1pc \hangafter=1 \no{Scoville, N.Z. 1972, ApJ 175, L127.} \hangindent=1pc \hangafter=1 \no Scoville, N.Z., Solomon, P. M., and Jefferts, K. B. 1974 ApJ 187, L63. \hangindent=1pc \hangafter=1 \no Sofue, Y. 1985 PASJ 37, 697 \hangindent=1pc \hangafter=1 \no Sofue, Y. 1989, in The Center of the Galaxy (IAU Symp. 136), ed. M.Morris (D.Reidel Publ. Co., Dordrecht) p. 213. \hangindent=1pc \hangafter=1 \no Sofue, Y. 1990 PASJ 42, 827 \hangindent=1pc \hangafter=1 \no Sofue, Y. 1991 PASJ 43, 671 \hangindent=1pc \hangafter=1 \no{Sofue, Y., and Handa, T. 1984, Nature 310, 568.} \hangindent=1pc \hangafter=1 \no Sofue, Y., Nakai, N. 1993 PASJ 45, 139. \hangindent=1pc \hangafter=1 \no Sofue, Y., Nakai, N. 1994 PASJ 46, 147 \hangindent=1pc \hangafter=1 \no{Sofue, Y., and Reich, W. 1979, AA Suppl 38, 251. \hangindent=1pc \hangafter=1 \no{Solomon, P.M., Scoville, N.Z., and Sanders, D.B., 1979, ApJ 232, L89.} \hangindent=1pc \hangafter=1 \no{S$\phi$rensen, S. -A., Matsuda, T., and Fujimoto, M. 1976, A. Sp. Sci. 43, 491. } \hangindent=1pc \hangafter=1 \no Stark, A. A., Bally, J., Wilson, R. W., Pound, M. W., 1989, in The Center of the Galaxy (IAU Symp. 136), ed. M.Morris (D.Reidel Publ. Co., Dordrecht) p. 213. \hangindent=1pc \hangafter=1 \no Tsuboi, M. 1989 in The Galactic Center (IAU Symp. 136), ed. M.Morris (D.Reidel Publ. Co., Dordrecht) p. 135 \hangindent=1pc \hangafter=1 \no Wada, K., Habe, A., Taniguchi, Y., Hasegawa, T. 1994, submitted to Nature \hangindent=1pc \hangafter=1 \no Wada, K., Habe, A. 1992 MNRAS 258, 82 \vfil\break \hsize=180truemm \vsize=250truemm \settabs 7 \columns \noindent Table 1: Galactic Center Arms and Ring. \vskip 2mm \hrule \vskip 0.4mm \hrule \vskip 2mm \+ Parameters & & Ring (I+II)& Arm I & Arm II& Arm III& Arm IV \cr \vskip 2mm \hrule \vskip 2mm \+ From $(l, V)$&($^\circ$, km s$^{-1}$) \dotfill &(+0.9,90) & $(0.9,80)$ & $(0.1,60)$ &$(0,140)$ &$(0.1,60)$ \cr \+& & & & $\sim(1,100)$ & & \cr \+ To $(l, V)$&($^\circ$, km s$^{-1}$) \dotfill & $(-0.65,-140)$ &$(-0.7, -150)$ &$(-0.6,-80)$ &$(-0.15,10)$ & $(0,-20)$ \cr \+& & & $\sim(-1,-200)$ & & & \cr \+ $V_{\rm LSR}$ at $l=0^\circ$&(km s$^{-1}$) \dotfill &~~~~.... &$-40$ &+50 &+70 & $-50$ \cr \vskip 2mm \+ From $(l, b)$&($^\circ$,$^\circ$)\dotfill &$(+0.9,0.0)$ &$(+0.9,-0.1)$ & $(0.25,-0.05)$ & $(0.25,0.25)$ &~~~~.... \cr \+ To $(l, b)$&($^\circ$,$^\circ$) \dotfill & $(-0.65,-0.08)$& ~$(-1.0,-0.2)$ &$(-0.65,-0.17)$ &$~(0,0)$ & ~~~~....\cr \+ $b$ at $l=0^\circ$&($^\circ$) \dotfill &~~~~.... & 0.050& $-0.067$&$(0,0)$ &~~~~.... \cr \+ Length &($^\circ$/pc) \dotfill &~~~~.... & 1.9/280 & 0.9/133 & 0.35/52&~~~~....\cr \+ Min. $b$ width&($^\circ$/pc)\dotfill &~~~~.... &0.088/13 & 0.091/13.5&~~~~.... &~~~~....\cr \+ Max. $b$ width&($^\circ$/pc)\dotfill &~~~~....& 0.33/50 & 0.2/30&~~~~....&~~~~.... \cr \v\v \+ Maj.ax.len.&($^\circ$/pc)\dotfill &1.55/230 & ~~~~.... &~~~~.... & {}~~~~....&~~~~.... \cr \+ Min.ax.len.&($^\circ$/pc)\dotfill &0.132/19.5 & ~~~~.... &~~~~.... &~~~~.... &~~~~.... \cr \+ Inclination&($^\circ$)\dotfill &$85^\circ.1$ & ~~~~.... &~~~~.... &~~~~.... &~~~~.... \cr \+ Ring cen. $(l, b)$& ~~($^\circ$,$^\circ$) \dotfill &(0.12,0.0) & ~~~~.... &~~~~.... &~~~~.... &~~~~.... \cr \+ Ring radius & (pc) \dotfill & 120 & & & & \cr \+ Rot. Velo &(km s$^{-1}$)\dotfill &$+90/-140$ & ~~~~.... &~~~~.... &~~~~.... &~~~~.... \cr \v\v \+ Mol. Mass$^\dagger$& ($10^7M_{\odot \hskip-5.2pt \bullet}$)\dotfill & 3.07 & 1.72 & 1.35 & {}~~~~....& ~~~~.... \cr \+ Remarks & \dotfill & Circum Nuc.& asso. Sgr B & Sgr C & Sgr A? & Sgr A?\cr \vskip 2mm \hrule \vskip 2mm \noindent $*$ The distance to the galactic center is assumed to be 8.5 kpc. \noindent $\dagger$ 1.61 times the H$_2$\ mass obtained from the $^{13}$CO intensity to H$_2$\ conversion [see eq. (1)-(3)], where the metal abundance has been assumed to be twice the solar. This also applies to mass in Table 2. The statistical error which occurs during intensity integration is only a few \%, while the error arising from ambiguity of the conversion factor is about 20 to 30\% (Arimoto et al 1994). \vfil\break \section{Figure Captions} Fig. 1: (a) The $(l, V)$\ diagram of the $^{13}$CO $(J=1-0)$ line emission of the central region of the Milky Way by averaging the data from $-0.35 \le b \le 0^\circ.17$ as obtained with the Bell Telephone 7-m telescope by Bally et al (1987). Contours are in unit of K $T^*_{\rm A}$\ at levels $0.1\times$(1, 2, 3, 4, 5, 6, 8, 12, 15, 20, 25). (b) The same as Fig. 1a, but the local and foreground CO emissions have been subtracted by applying the ``pressing method'' (see the text for the procedure). Contour levels are same as in (a). \v\v Fig. 2: The $(l, V)$\ diagrams averaged in $4'~b$ interval. Local/foreground emissions have been removed. `Rigid-rotation' ridges (arms) are dominant in the disk at $|b|<\sim 10'$ (25 pc). Contours are in unit of K $T^*_{\rm A}$\ at levels $0.2\times$(1, 2, 3,..., 9, 10, 12, 14, 16, 18, 20, 25, 30). \v\v Fig. 3: Schematic sketch of the major ridges (arms) in the $(l, V)$\ diagrams. \v\v Fig. 4: (a) Integrated intensity map in the whole velocity range at $-250 \le V_{\rm LSR} \le 250$ km s$^{-1}$. This is almost the same as the map presented by Stark et al (1989), except that the local contribution has been subtracted. Contours are in unit of K km s$^{-1}$\ at levels $25 \times$(1, 2, 3, ..., 9, 10, 12, 14, 16, 18, 20, 25, 30). (b) Same but in a grey-scale representation. For intensity scale, see (a). The bottom figure shows the same, but the scale in the latitude direction has been doubled. Galactic Center Arm (GCA) I runs as a long arc in the positive $b$ side; GCA II runs in the negative $b$ side. \v\v Fig. 5: $(l, V)$\ diagrams corresponding to (a) Galactic Center Arms I and (b) II, which were used to obtain intensity maps of the Galactic Center Arms in Fig. 6. Contours are in unit of K $T^*_{\rm A}$\ at levels $0.2\times$(1, 2, 3, 4, 5, 6, 8, 10, 12, 15, 20, 25, 30, 35). \v\v Fig. 6: Integrated intensity maps corresponding to (a) Galactic Center Arm I, and (b) Arm II as in Fig. 5. Contours are in unit of K km s$^{-1}$\ at levels $12.5\times$(1, 2, 3, ..., 9, 10, 12, 14, 16, 18, 20, 25, 30). (c) Arms I+II. Contours are in unit of K km s$^{-1}$\ at levels $25\times$(as above). \v\v Fig. 7: Intensity variation across Galactic Center Arms I and II perpendicular to the galactic plane averaged at $l=0^\circ.24$ to $-0^\circ.33$, where the arms are most clearly separated. \v\v Fig. 8: (a) A velocity field as obtained by taking the first moment of the $(V_{\rm LSR}, l, b)$ cube (intensity-weighted mean velocity field). Contour interval is 10 km s$^{-1}$\. Full-line contours are for positive velocity starting at 0 km s$^{-1}$. Dashed contours are for negative velocity. (b) Same as (a), but for the ``disk component'' with $|V_{\rm LSR}|<100$ km s$^{-1}$. \v\v Fig. 9: Two-armed spiral model with a spiral infalling motion. Gas density distribution is shown by spiral-like contours as projected on the galactic plane. Calculated $(l, V)$\ diagram is shown by tilted X shaped contours. The scales are arbitrary. (a) Two spiral arms with a pitch angle $p=10^\circ$ are assumed. The azimuthally averaged gas density has a hole at the center, corresponding to a ring distribution of gas on which two arms are superposed. In addition to a constant circular rotation, radial infall of velocity $V_{\rm rot} {\rm sin}~p$ is superposed. (b), (c) The same as (a), but the spiral arms are oval in shape whose major axis are inclined by $\pm 30^\circ$ from the nodal line. (d)-(f) The same as (a)-(c), respectively, but the density distribution along the arms has the maximum at the center and the pitch angle is taken larger: $p=20^\circ$. \v\v Fig. 10: Possible deconvolution of the $(l, V)$\ diagrams for Galactic Center Arms I and II into a spatial distribution as projected on the galactic plane. Contour interval is 0.25 starting at 0.1 in an arbitrary unit. Sgr A is assumed to be at the center. \v\v Fig. 11: Superposition of the radio continuum emission at 10 GHz (contours: Handa et al 1987) on (a) $^{13}$CO, and (b) CS emission maps (grey scale). Contours are in unit of K $T_{\rm B}$\ of 10 GHz continuum brightness at levels $0.1\times$(1, 2, 3, 4, 6, 8, 10, 15, 20, 25). For CO intensity scale, see Fig. 4. \bye
\section{Varieties of Factorization and Evolution} With the extraordinary data from the Tevatron of the last runs, our knowledge of large momentum transfer processes has taken a qualitative step forward. The demands on theory are now much more stringent, and afford tests of QCD concepts that were not previously practical. In particular, TeV energies allow for many two-scale processes where both scales are much larger than a GeV, and hence may be amenable to perturbative treatment. For instance, $\alpha_s(t)\; \ln(s/t)$ may be a relatively large number even for $\alpha_s(t)$ small, necessitating a resummation of powers $(\alpha_s(t)\; \ln^2(s/t))$ and $(\alpha_s(t)\; \ln(s/t))$ to all orders in perturbation theory. In the following, I shall review the basic techniques used to study two-scale inclusive cross sections at low parton density. For such process factorization in terms of single-parton densities holds, \begin{equation} \sigma_H(p,q) = \sum_i \int dx' {\hat \sigma}_i(x'p,q)\; \phi_{i/H}(x')\, , \label{factor} \end{equation} with ${\hat \sigma}_i$ a short-distance partonic cross section and $\phi_{i/H}$ the distribution of parton $i$ in hadron $H$. I will begin with a review of three basic evolution equations, which govern the behavior of parton distributions in three limits of particular physical interest, and which are commonly referred to as the DGLAP, BFKL and Sudakov equations. Of these, the DGLAP is the best-studied, and controls the standard evolution of parton distributions in $Q^2$ for values of $x$ for which neither $\alpha_s\ln x$ nor $\alpha_s \ln(1-x)$ is a large number. The BFKL equation governs the behavior of parton distributions at small $x$ and fixed $Q^2$ and of amplitudes near the forward direction. Finally, the Sudakov evolution equation describes the behavior of distributions and amplitudes in the elastic limit ($x\rightarrow 1$ for deep-inelastic scattering). These limits are illustrated schematically in Fig.\ 1. \begin{figure}[ht] \centerline{\epsffile{pbpfig1d.eps}} \caption{Actions of evolution equations.} \label{fig1} \end{figure} Each of these evolution equations is associated with the ladder structure shown in Fig.\ 2, in which parton $j$, with momentum $p$ ``evolves" into parton $i$, of momentum $xp$, in the process emitting $N$ quanta. In perturbation theory, the distribution of parton $i$ in parton $j$ may be thought of as a sum over $N$, \begin{equation} \phi_{i/j}(x,q_T) \sim \sum_N \phi_{i/j}^{(N)}\, . \label{laddersum} \end{equation} \begin{figure}[ht] \centerline{\epsffile{pbpfig2c.eps}} \caption{General ladder structure.} \label{fig2} \end{figure} \subsection{DGLAP} The basic evolution equation for perturbative QCD, originated by Dokshitser, Gribov, Lipatov, Altarelli and Parisi \cite{DGLAP}, has its origin in the description of deeply inelastic scattering (DIS). The factorization theorem for any DIS structure function $F$ in hadron $H$ is \begin{equation} F^{(H)}(x,Q^2) = \int_x^1 {dx'\over x'}\; C_{j}(x/x')\; \phi_{j/H}(x,Q^2)\, , \label{DISfact} \end{equation} with perturbative short-distance functions $C_i$ and nonperturbative parton distribution(s) $\phi$. Although $\phi$ is not perturbatively calculable, its $Q^2$ dependence is, in terms of kernels, or splitting functions, $P_{ij}$, through the DGLAP evolution equation \begin{equation} { d\over d \ln Q^2}\phi_{i/H}(x,Q^2) = \sum_{j=f,{\bar f},G} \int_x^1 {d \xi \over \xi}\; P_{ij}(x/\xi)\; \phi_{j/H}(\xi,Q^2)\, . \label{dglap} \end{equation} The universality of the distributions allow us to derive predictions for hadronic collisions from parton distributions derived (primarily) from DIS. The $P_{ij}$ themselves are known up to two loops. In general, the DGLAP equation is applicable when the evolution is generated by the emission of quanta with strongly ordered transverse momenta ($k_{1T}\ll k_{2T} \ll \dots \ll k_{NT}$ in Fig.\ 2, for instance). The integrals over these transverse momenta give logarithmic enhancements, which are organized by the DGLAP equation. There are other sources of logarithmic enhancement, however, already present as factors of $1/x$ in the splitting function $P_{GG}(x)$. Such singular behavior produces logarithms of $x$, even in the absence of large transverse momentum enhancements. The BFKL equation, to which we now turn, summarizes the effects of these logs of $x$. \subsection{BFKL} The BFKL equation (Balitskii, Fadin, Kuraev and Lipatov \cite{BFKL}) was developed originally to resum logs of $s$ in the total hadronic cross section (see below). We begin our consideration of it here, however, with a generalization of DIS factorization (\ref{DISfact}), to a form that links the distributions in both longitudinal and transverse momenta \cite{ktfactci}, \begin{equation} F(x,Q^2)=\int_x^1 {dx' \over x'}\; d^2k_T\; C(x/x',{\bf k},Q^2)\, {\cal F}(x',{\bf k})\, , \label{ktfact} \end{equation} with $C(\xi,{\bf k},Q^2)$ a new, transverse momentum-dependent short-distance function and $\cal F$ the corresponding parton distribution. In general, the gluon distribution tends to dominate at low $x$, and we shall restrict our discussion to ${\cal F}_G$. Eq.\ (\ref{ktfact}) is of particular interest when $x\rightarrow 0$, so that the mass of the final state in DIS, $(1-x)Q^2/x$, grows. The BFKL equation, which describes the behavior of ${\cal F}_G$, may be written in many forms, one of the least intimidating of which is \begin{equation} {d\over d \ln x}{\cal F}_G(x,{\bf k}) = {3\alpha_s\over \pi} \int{d^2{\bf k}' \over ({\bf k}'-{\bf k}')^2}\ \left ( K\; *\; {\cal F}_G(x,{\bf k}') \right )\, , \label{bfkl} \end{equation} with a kernel defined by \begin{eqnarray} K\; *\; {\cal F}_G(x,{\bf k}') &=& {\cal F}_G(x,{\bf k}') \nonumber \\ &\ & \quad - \left ( {{\bf k}\cdot{\bf k}'\over 2{{\bf k}'}^2} \right ) {\cal F}_G(x,{\bf k})\, . \label{bfklkernel} \end{eqnarray} An excellent introduction to the BFKL equation may be found in the recent lectures of Del Duca \cite{DelDuca}. The BFKL equation is somewhat harder to solve than the DGLAP equation, but it has many applications, including DIS for small $x$, semihard processes (minijet, heavy quark, etc.) in hadronic collisions, and color singlet exchange at $|t|\ll s$. Although the BFKL formalism has traditionally remained at leading logarithm in $x$, recent progress has been reported on generalizations that determine a two-loop kernel \cite{BFKL2loop}. In addition, it is possible to develop a generalized evolution equation that interpolates between the DGLAP and BFKL equations at leading logarithm \cite{Marches}. For this purpose, the angle of emission for soft gluons may be chosen as the primary evolution variable, reflecting an ordering of angles in sequential emission \cite{Muellangord}. Control of such ``coherence" effects in soft QCD radiation \cite{cohere} also serves as an important ingredient in the construction of detailed models of QCD final states \cite{herwig}. \subsection{Sudakov} The other fundamental limit of evolution is $x\rightarrow 1$. In DIS, this corresponds to a low-mass final state with high energy. This ``elastic" limit is illustrated in Fig.\ 3. In addition to soft radiation, a well-collimated jet of mass near $(1-x)Q^2$ emerges from the hard scattering $H$. \begin{figure}[ht] \centerline{\epsffile{pbpfig3.eps}} \caption{DIS as $x\rightarrow 1$.} \label{fig3} \end{figure} Corresponding to this physical situation, we may identify a factorized form for such a cross section, \begin{equation} F(x,Q^2) = H(Q)\; \int_x^1 dx'\; J((1-x')Q^2,Q)\, \phi(x',Q^2)\, , \label{Sudfact} \end{equation} with $J$ a universal function representing the hadronization of the jet (in this form certain non-leading logarithmic factors are suppressed). For $(1-x)Q^2\gg \Lambda^2$, $J$ is computable in perturbation theory. Given the factorization (\ref{Sudfact}), the $x$ dependence of the jet may be determined from the Sudakov evolution equation \cite{St87}, \begin{eqnarray} \left ( {\partial \over \partial \ln(1-x)} + {1\over 2}\beta(g){\partial \over \partial g}\; \right ) J((&1&-x')Q^2,Q) = {1\over 2}S(\alpha_s)\; J((1-x')Q^2,Q) \nonumber \\ &\, & -{1\over 2}\int_x^1 {d y\over y}\; K_J(1-x/y)\; J((1-y)Q^2,Q)\, , \label{Sudevol} \end{eqnarray} whose kernel $K_J$ is of the form of a plus distribution, \begin{equation} K_J(1-z) = \left [ {\Gamma_J(\alpha_s((1-z)^2Q^2)\over 1-z} \right ]_+\, . \label{Sudkernel} \end{equation} The function $\Gamma_J$ is a universal Sudakov anomalous dimension \cite{CoSo81}, related to the $\ln n$ dependence of the standard DGLAP anomalous dimensions $\gamma_n$ \cite{KoTr}. Thus, $\Gamma_J$ is also known to two loops. $S$ is a power series in $\alpha_s(Q^2)$. Sudakov evolution finds applications to jet event shapes, threshold corrections and transverse momentum distributions, some of which we shall review below. \section{Resummation for Small $x$ and Diffractive Cross Sections} In this section, I will review a few of the prominent applications of the BFKL resummation of small-$x$ enhancements, beginning with DIS. \subsection{DIS} For DIS, the object of interest is the $k_T$-dependent gluon distribution, the solution to eq.\ (\ref{bfkl}). Eigenfunctions of the derivative may be found by direct substitution. The dominant power as $x\rightarrow 0$ specifies the growth of the gluon distribution. It is \cite{BFKL,DelDuca} \begin{equation} {\cal F}_G(x,{\bf k}) \sim x^{-\omega_0}\, , \label{bfklsoln} \end{equation} times a power of ${\bf k}^2$, where \begin{equation} \omega_0={4N\alpha_s \over \pi}\ln 2\, , \label{omegao} \end{equation} with $N=3$ the number of colors. Since quark distributions mix with the gluon distribution, BFKL evolution suggests that the structure function of any hadron behaves as \begin{eqnarray} F_2(x) &\sim& \sum_i Q_i^2\; x\; \phi_i(x) \nonumber \\ &\rightarrow& x^{-\omega_0}\, , \label{ftwopom} \end{eqnarray} which diverges as a power. This behavior is referred to as that of the ``bare" pomeron. It is clear that such growth cannot be supported to arbitrarily small $x$, since if parton densities grow too large, they will begin to interfere or ``shadow" each other. In technical terms, the assumption of low-density partons, which underlies each of the evolution equations above, fails for $x$ low enough \cite{shadow}. HERA \cite{zeusH1} has seen a growth in DIS structure functions at small $x$ which is consistent with the qualitative expectations of BFKL resummation. The DGLAP equation, however, based on $k_T$-ordering, also results in growth for $x$ small, if not quite so dramatic. Untangling the physical content of the HERA data is a subject of great current interest \cite{smallxtheo}. \subsection{BFKL in Hadron-hadron Scattering} The BFKL formalism \cite{BFKL} was originally developed to describe hadron-hadron scattering in QCD, both the total cross section and the closely-related Regge limit, $t$ fixed, $s\rightarrow \infty$. Its basic consequences for inclusive hadron-hadron scattering near the forward direction may by summarized by the following ``translation" from DIS notation, \begin{eqnarray} M^2_{\rm had} = {1-x \over x}Q^2 &\rightarrow& s \nonumber \\ x^{-\omega_0} &\rightarrow& s^{\omega_0} \nonumber \\ F(x) &\rightarrow& \sigma_{\rm tot} \sim s^{\omega_0}\, . \label{distohh} \end{eqnarray} Note in particular that the total cross section grows as a power, corresponding to the $x\rightarrow 0$ behavior of $F_2(x)$ in DIS. Again, a rise in the total cross section for hadron-hadron scattering has long been seen at high energy, but an uninterrupted power-law rise would violate unitarity, as embodied in the Froissart bound. The ``bare" pomeron identified above therefore cannot be the final answer. I shall return to recent progress on this question below. It is of some interest to sketch the momentum-space configurations that give rise to the BFKL pomeron in hadron-hadron scattering. These are illustrated by the ladder diagrams in Fig.\ 4, where the vertices $C_i$ summarize contributions not only from the diagrams explicitly shown, but also from non-ladder diagrams of the same order that are important in gauge theories \cite{BFKL}. \begin{figure}[ht] \centerline{\epsffile{pbpfig4a.eps}} \caption{BFKL ladders.} \label{fig4} \end{figure} These vertices are related to the kernel in eq.\ (\ref{bfkl}) by $C_i^2\sim K$. As above, all lines in the ladder represent gluons. The growth of the cross section results from the production of many particles, whose momenta are conveniently parameterized in components parallel to the incoming momenta $k_a$ and $k_b$, and transverse components, \begin{equation} k_i=\alpha_ik_a+\beta_ik_b+k_{iT}\, . \label{ksubi} \end{equation} The BFKL pomeron resums logarithms that result from configurations in which the rapidities $y_i\sim \ln (\alpha_i/\beta_i)$, are strongly ordered, but the transverse momenta are all of the same order as the momentum transfer, \begin{eqnarray} \alpha_1 \gg \alpha_2 &\gg& \cdots \gg \alpha_n \nonumber \\ \beta_1 \ll \beta_2 &\ll& \cdots \ll \beta_n \nonumber \\ k_{iT} &\sim& k_{jT}\, . \label{strongorder} \end{eqnarray} It is the lack of ordering in transverse momentum that distinguishes BFKL evolution. Studies show that kinematic signs of this evolution should be present in final states \cite{bfklkinem}, and data from the Tevatron and HERA are being closely scrutinized for evidence of these effects. \subsection{Soft and Hard Diffraction} Ampitudes for soft and hard diffraction are illustrated in Fig.\ 5. In 5a, the lines in both the cut and uncut ladder are ordered in rapidity, just as in Fig.\ 4. \begin{figure}[ht] \centerline{\epsffile{pbpfig5b.eps}} \caption{Schematic representations of soft (a) and hard (b) diffraction.} \label{fig5} \end{figure} Only the cut lines, closer in rapidity to the incoming top line appear in the final state, however, leading to a ``gap" in rapidity, and an apparent ``diffractive excitation" of the top line. The uncut ladder, a virtual pomeron, acts very much like a virtual elementary particle. Such events have been seen for some time, with particles that emerge into the final state generally at low transverse momentum. More recently, however, events have been seen that have both rapidity gaps and high momentum transfers \cite{rapgap}. A possible mechanism is illustrated in Fig.\ 5b, in which the incoming top line undergoes a hard scattering from a ``parton" of the pomeron. This idea \cite{IngSch} suggests that it might make sense to talk about the distributions of partons in pomerons. Such a formalism would be useful if these cross sections enjoy the universality and incoherence properties that underly normal factorization for inclusive cross sections \cite{Collinspomphi}. Beyond this, the concept of rapidity gaps as a sign of color singlet exchange, even at large momentum transfer \cite{singletrapgap} is an attractive concept. As noted above, the BFKL formalism treats leading logarihms of $s$. For singlest exchange, these are single logarithms per loop, $\alpha_s^n\ln^ns$. More generally, elastic scattering of, say, high-energy quarks may involve both singlet and octet exchange \begin{equation} A = A_8T_8 + A_1 T_1\, , \label{coloramp} \end{equation} with the $T_i$ tensors in the color space. Supposing that $t$ is large, but $|t|\ll s$, it is possible to study the leading logarithms assoicated with soft gluon exchange according to Sudakov evolution methods \cite{softSud}. The results for the $A_i$, neglecting for simplicity the effects of the running coupling, are of the schematic form \cite{K,StS} \begin{equation} A_8\sim e^{-\ln s\ln t}\, \, ,\, \, A_1 \sim e^{-\ln^2 t}\, . \label{colorsud} \end{equation} This is consistent with BFKL resummation of logs of $s$ in singlet exchange at fixed $t$. For octet exchange, it is equivalent to a ``Reggeized" gluon behavior \cite{BFKL}. These results are also consistent with the observation of rapidity gaps in hard scattering at high energy, and with their connection to singlet exchange. They are also relevant to a very different process, near-forward hard elastic scattering \cite{LandDonn}, in which an observed behavior of $t^{-8}$ may be a sign of triple pomeron exchange between valence quarks \cite{StS}. \subsection{Taming the Bare Pomeron} As noted above, the bare pomeron violates the Froissart bound. Considerable effort recently has gone into remedying this defect, and to creating a formalism that reduces to the BFKL analysis for moderate $s$, but which has built-in unitarity. Here, I shall mention only a few examples. Lipatov has developed such a scheme, based on adding a subset of nonleading logarithms in $s$ and employing an expansion in $1/N$, with $N$ the number of colors. In this scheme, the pomeron is a sort of bound state of Reggeized gluons \cite{lipconf}. Adding more such gluons, the problem reduces to finding the spectrum of a two-dimensional field theory in impact parameter space. Recently, Faddeev and Korchemsky \cite{FadKor} have shown that the resulting model is ``solvable" in the sense of possessing infinite numbers of conservation laws. They do not, however, explicitly ``solve" it. Another approach toward deciphering the BFKL pomeron involves studying a model in which it arises self-consistently in perturbation theory, the scattering of bound states of heavy quarks \cite{qqbarbfkl}. Although not yet realistic, such models allows a quite detailed study of soft gluon dynamics at large $N$ and leading logarithm. \section{Resummation at the edge of phase space} I now turn to applications of Sudakov resummation, beginning with a typical example from event shapes, the thrust. \subsection{Thrust} The thrust in ${\rm e}^+{\rm e}^-$ annihilation is defined by \begin{equation} T= {\rm max}_{\hat n} {1\over Q} \sum_i\, |{\vec p}_i\cdot {\hat n}|\, , \label{thrust} \end{equation} with the sum over all particles in the final state, and with a maximum taken over all unit vectors $\vec n$. In perturbation theory, the fixed-thrust cross section diverges at $T=1$, which is the elastic limit of two lightlike back-to-back jets. A measure of this divergence is given at one loop by \begin{equation} {1\over \sigma_0}\, \int_0^T dT' {d\sigma \over dT'} \sim 1- {\alpha_s\over \pi}C_F\; \ln^2(1-T)\, {}. \label{oneloopthrust} \end{equation} If the two jets have masses $p_i^2\sim (1/2)(1-x_i)Q^2$, then near the edge of phase space $x_1=x_2=1$, we have \begin{equation} T \sim {1\over 2}\left [ (1-x_1)+(1-x_2)\right ]\, . \label{twojetthrust} \end{equation} In the same region, however, the cross section (simplified by neglecting certain nonleading corrections) factors into the product of the jet functions described above, \begin{eqnarray} {1\over \sigma_0}\, {d\sigma \over dT'} &\sim& \int dx_1dx_2\; \delta\left(1-T-{1\over 2}(2-x_1-x_2)\right ) \nonumber \\ &\ & \quad\quad\quad \times J_1((1-x_1)Q^2,Q)\; J_2((1-x_2)Q^2,Q)\, , \label{thrustfact} \end{eqnarray} each of which satisfies a Sudakov evolution equation, whose solution gives \begin{eqnarray} \int dx\, e^{-\nu(1-x)Q^2/2}\; &J&((1-x)Q^2,Q) \nonumber \\ &\sim& \exp \left [ -\int_0^1 {du\over u}(1-e^{-u\nu}) \int_{u^2Q^2}^{uQ^2} {d\mu^2\over \mu^2} \Gamma_J(\alpha_s(\mu^2)) \right ]\, . \label{thrustresum} \end{eqnarray} This result \cite{thrustresum}, and its analog for the transverse momentum distribution in Drell-Yan and Z$_0$ production \cite{qtresum}, are very helpful in understanding high energy data. \subsection{Threshold Resummation} Another important application of Sudakov resummation is to high-mass Drell-Yan, top and other cross sections where the parton distributions are falling rapidly at threshold. The simplest example is the inclusive Drell-Yan cross section, \begin{equation} {d\sigma \over dQ^2} = \sigma_0\, \sum_q\, \int_\tau^1 dz\; \Phi_{q{\bar q}}(z)\; \omega_{q{\bar q}}(z,\alpha_s(Q^2))\, , \label{DYfact} \end{equation} where the partonic flux is given by \begin{equation} \Phi_{q{\bar q}}(z) = \int_0^1 {dx_adx_b\over x_ax_b}\; \delta\left(1-\tau/(zx_ax_b)\right)\; Q_q^2\; \phi_q \left (x_a,Q^2\right ) \phi_{\bar q} \left (x_b,Q^2\right )\, , \label{Phitauz} \end{equation} with $Q_q$ the quark charge in units of electron charge and with $\sigma_0=4\pi\alpha^2/3NQ^2s$ the Born cross section for $Q_q=1$. Here the overall partonic invariant is ${\hat s}=Q^2/z$, with threshold at $z=1$. Specializing to DIS scheme, initial state radiation from the incoming quark pair in Drell-Yan cancels against initial state radiation in the DIS structure function, illustrated in Fig.\ 3 above. The (infrared safe) radiation from the outgoing quark jet in DIS, however, is not compensated. But since, as indicated above, the $x\rightarrow 1$ behavior of this jet can be computed, we can give an explicit form for the hard-scattering cross section $\omega_{q{\bar q}}$ in eq.\ (\ref{DYfact}) that includes all $\ln^n(1-x)$ behavior in this limit. Schematically, it is of the form \begin{equation} \omega_{q{\bar q}}(z,\alpha_s(Q^2)) = \delta(1-z) - \left [ {\rm e}^E(1-z,\alpha_s) F(1-z) \right ]_+\, , \label{omegaresum} \end{equation} where the function $F(1-z)$ is given by \cite{CoSt} \begin{equation} F(1-z)={1 \over \pi}\; \sin(\pi P_1)\Gamma(1+P_1) +\dots\, , \label{Fdef} \end{equation} with $P_1$ the derivative of the exponent $E$ with respect to $\ln (1-z)$, \begin{equation} P_1(1-z,\alpha_s)={d \over d\ln(1-z)}E(1-z,\alpha_s)\, . \label{ponedef} \end{equation} $E(1-z,\alpha_s)$ is itself of the moment form, \begin{equation} E(n,\alpha_s) = \int_0^1 dx \left ({x^{n-1}-1 \over 1-x}\right ) \int_{(1-x)^2Q^2}^{(1-x)Q^2} {d\mu^2 \over \mu^2}\, g_1(\alpha_s(\mu^2)) + \dots \label{Edef} \end{equation} where $g_1$ is closely related to the jet anomalous dimension $\Gamma_J$, discussed above, \begin{equation} g_1(\alpha_s)=2C_F\; {\alpha_s \over \pi}+\dots=2\Gamma_J(\alpha_s)+\dots\, . \label{geeone} \end{equation} Given the form of $E$, we expect the effect of these corrections to be positive, and to raise the cross section above low order predictions. This expectation has been borne out by explicit calculations in leading-logarithm approximation for the DY process \cite{AMS}. We have heard elsewhere at this conference that one-loop QCD predictions fall short of the very high-$E_T$ jet cross section measured by CDF. This cross section is of the same general form as above, given schematically by \begin{equation} { d\sigma_{1J} \over dE_T}\bigg |_{y=0} = \int_{2E_T/\sqrt{s}}^1\; \Phi(z)\; { d{\hat \sigma}_{1J} \over dE_T} ({\hat s}=zs)\; dz\, . \label{jetsigma} \end{equation} It is natural to ask whether this effect might be due to resummable soft gluon effects. Now we must note that in jet production final-state as well as initial state interactions may be important. Nevertheless, final-state interactions cancel in any cross section with finite angular and energy resolutions, leaving the same, universal initial-state corrections as in DY. It will be of great interest to test these ideas in the coming months, since an excess of high-energy jets relative to QCD predictions could indicate profound new physics. \subsection{Regulated Resummation} Eq.\ (\ref{Edef}) for the resummed exponent has a problem at small $1-x$, where the perturbative running coupling may diverge. To make sense of (\ref{Edef}), one may use a cutoff in $x$ or $k_T^2$. Another possibility, following the example of ref.\ \cite{Mu85}, is to define the integral as a principal value \cite{CoSt2}. There is no special status for such a definition, but it enables us to express $E$ explicitly in terms of special functions, to be specific as a series of exponential integrals. In fact, any such definition of the perturbative expansion can be made only at the price of adding new nonperturbative parameters to the theory, as \begin{equation} E_{NP} = \sum_{n\ge 1} {A_n \over Q^n}\, . \label{enp} \end{equation} Cutoff integrals are probably numerically simpler. A sample application to resummed leading logarithms was studied in \cite{AMS} for the Drell-Yan process, \begin{equation} \int_0^1 dx\; {x^{n-1} -1 \over 1-x} \rightarrow \int_0^{1-\lambda/\Lambda} dx\; {x^{n-1} -1 \over 1-x}\, . \label{cutoff} \end{equation} Order by order, dependence on $\lambda$ is higher twist. For large enough $Q^2$, a cutoff resummation of this form is relatively insensitive to $\lambda$, but grows with $\tau$ at fixed $Q^2$ \cite{AMS}. Studies of the behavior at fixed $s$ are currently underway. Finally, we may note that a cutoff prescription was employed in \cite{LSvN} to define resummed perturbation theory for top production, with partonic cross sections of the form \begin{equation} \sigma_{ij}(z,m_t^2,s_0) = \int_{s_0}^{s-2m_t\sqrt{s}} ds_4\; e^{{\bar E}(s_4)}\; {d\sigma^{(0)}_{ij} \over ds_4}\, , \label{lsvn} \end{equation} where $s_4=0$ at threshold, where ${\bar E}(s_4)$ is a resummed exponent related to (\ref{Edef}) and where $\sigma^{(0)}_{ij}$ is the Born cross section. \section{Infrared Renormalons and a New Source of Power Corrections} Expanding the running coupling integrand in eq.\ (\ref{Edef}) in terms of $\alpha_s(Q^2)$, we derive integrals of the form \begin{equation} (-b_2)^m\alpha_s^m\; \int_0^{Q^2} {dk_T^2\over Q^2}\;\ln^m \left({k_T^2\over Q^2}\right ) =b_2^m\alpha_s^m\; m!\, , \label{irrenorm} \end{equation} in which the singularity in the running coupling is reflected in factorial growth of expansion coefficients in perturbation theory. This phenomenon is commonly referred to as an ``infrared renormalon" \cite{tHooft}. A closer analysis shows that such behavior may be interpreted as an ambiguity in the perturbative expansion, which may be removed by modifying the expansion to make it convergent, while at the same time adding a new term of the form of eq.\ (\ref{enp}), $A_N/Q^N$, with $Q$ the momentum transfer. The classic case \cite{Mu85} is the total cross section in ${\rm e}^+{\rm e}^-$ annihilation, where the nonperturbative term is of the form $A_4/Q^4$, with $A_4=\langle F^2\rangle_0$, the vacuum expectation value of the squared gluon field strength. A similar analysis for resummed jet shapes \cite{StArg} and Drell-Yan cross sections \cite{CoSt2} shows that their resummed exponents potentially have a much stronger nonperturbative behavior \cite{W,KS}, \begin{equation} E=E_{\rm PT}^{({\rm reg})}+{\Lambda \over \delta Q}A\, , \label{interp} \end{equation} where $\delta=0$ in the elastic limit (infinitely collimated jets, for instance). As in the case of the total ${\rm e}^+{\rm e}^-$ cross section, it is possible to find a field-theoretic analog to $A$, but in this case it is in terms of a field-strength, integrated over the classical paths of the relevant partons. For example, for a two-jet cross section, with directions are $p_1$ and $p_2$, we find \cite{KS} \begin{eqnarray} \Lambda A &=& \langle 0|{\bar {\rm T}} \left ( \Phi^\dagger_{p_1}(0,\infty) ({\cal F}_{0p_1}^\dagger(0) -{\cal F}_{0p_2}(0,\infty)) \Phi_{p_2}(0,\infty) \right ) \nonumber \\ &\ & \quad\quad\quad \times {\rm T} \left (\Phi_{p_1}^\dagger(0,\infty) \Phi_{p_2}(0,\infty) \right ) |0\rangle\, . \label{Aop} \end{eqnarray} This matrix element is gauge invariant, with the $\Phi$'s defined in terms of ``Wilson lines", and the operator $\cal F$ in terms of the field strength. The rather general form of the operator suggests that these new corrections, suppressed by a single power of the large momentum scale, may possess universality properties. The concept of universality in this context has already been given much study \cite{KS,DokWeb}, and its promise and limitations have been discussed in \cite{KSMord}. \section{Conclusions} Soft-gluon resummation allows varied applications of quantum field theory that are relevant to QCD at current energies. One goal of this program is to achieve a new level of precision by combining large corrections to all orders. It also affords potential insights into features of nonperturbative structure, through the high-order behavior of the perturbation series. Regarding specific applications, resummed threshold corrections to very high energy jet cross sections urgently need more study, especially because they are relevant to signs of new physics. Also, studies of $-t\ll s$ hard elastic and jet cross sections are attractive for their relevance to hard pomeron and Sudakov effects and to the valence-quark structure of the nucleon. Finally, infrared renormalons from resummation imply the existence a new class of nonperturbative parameters that may be measured in, for instance, jet shape analysis, and which may shed new light on the perturbative/nonperturbative interface in QCD. \smallskip It is a pleasure to thank Lyndon Alvero, Harry Contopanagos, Gregory Korchemsky and Jack Smith for many helpful conversations and explanations. This work was supported in part by the National Science Foundation under grant PHY9309888.
\section{Introduction} In recent years the understanding of the effects of disorder in the physics of metals, semiconductors and amorphous systems has made a tremendous progress. This vigorous development was motivated to a great extent by a thorough understanding of how strongly disorder effects determine the behavior of real physical systems. The success of the various analytical descriptions which have been considered, however, has always been determined by the relative simplicity to which approximations could be reduced, in order to keep the theory tractable analytically, without losing its capability to account for the most important physical aspects. One of the most successful approximations to match these requirements has been the coherent potential approximation CPA developed by Soven \cite{bi1}, Taylor \cite{bi2}, and extended by Leath \cite{bi7,bi8,bi9}, Velick{\'y} \cite{bi3,bi4,bi5,bi6} and many others (see Ref. \onlinecite{bi13} for a review). In contrast to many other approximations for the procedure of configurational averaging in disordered systems, the CPA is capable of interpolating correctly between the limits of weak \cite{bi10} and strong disorder \cite{bi12} as well as low and high impurity concentrations. Therefore it is also able to predict accurately the formation of impurity bound states turning into split-off impurity bands as the impurity concentration is increased. Extensions of the single site CPA to include scattering from clusters of impurities are still today the only existing analytic theories that allow for a calculation of the density of states in disordered systems such that a reasonable versimilitude is attained \cite{bi14}. The main difficulty with the CPA, is the relative complexity of the self consistent equations which have to be solved in more accurate extensions of the theory. In the present paper we propose a more differentiated analysis of two particle properties within the CPA building on the single site approximation (SSA) in which it was developed originally. Our novel treatment should be especially useful for applications to binary substitutional alloys in all regimes of disorder. Following an earlier paper by Velick{\'y} \cite{bi4}, in which the regular two particle theory within the CPA was first introduced and which will be referred to hereafter as {\bf I}, the theory of weighted single particle resolvents \cite{bi26} is extended to a properly weighted two particle theory. Weighting of a Green's function in this context means that through the application of appropriate operators to the unaveraged resolvent, restrictions are made on the type of alloy component of either or both of the sites on which the particle starts and terminates its motion. Upon averaging, this results in a statistical weight being attributed to the averaged unrestricted resolvent with possibly separate additive terms. On the average, therefore, contributions to a full two particle Green's function in terms of constituent components can be resolved. This allows for a better understanding of how several physical processes contribute to a cumulative behavior as the principal parameters of the system are varied and it can therefore be used in an analysis of further effects which differentiates between these components. We have structured the paper as follows: In section II the important features of the single particle CPA for diagonal disorder are recapitulated and relations which are important to the calculation of the corresponding two particle Green's functions are established. In section III we calculate the Fourier transforms of a class of weighted two particle functions which are kept as general as possible. For that reason only a representative choice of weightings are calculated explicitly, since other weights can be obtained in an analogous fashion. Section IV is devoted to possible applications of weighted two particle functions in linear response theory, and the peculiarities of two different classes of such functions are discussed in detail, which we have selected to cover a large range of conceivable applications. In the third part of section IV the behavior of one class of functions discussed before is examined in a split band limit (strong disorder). Section V is devoted to a numerical study of the splitting into several components of an interband absorption spectrum of a disordered alloy. Section VI in conclusion discusses the implications and possible further applications of the results obtained throughout this work. \section{Single particle properties} We consider a binary substitutional alloy $A_c B_{1-c}$ on a simple totally disordered monoatomic lattice on which each site is occupied by either an $A-$ or a $B$-component with probabilities $c$ and $1-c$, respectively. Since by symmetry it is only necessary to consider concentrations between $0 \leq c \leq 0.5$, we restrict our investigation to this region and define $A$ as the impurity and $B$ as the host component of the alloy. To simplify further calculations we consider diagonal disorder only, i.e. a Hamiltonian of the form \begin{equation} H = H_0 + U = H_0 + \sum_n U_n \label{a0} \end{equation} where $H_0$ represents the periodic part and the $U_n $ are single site contributions to a random disorder potential which assume either the value $U^A_n = V$ or $U^B_n = 0$ with probabilities $c$ and $1-c$, depending on whether the site $n$ is an impurity or part of the host material. In principle one could also have introduced a symmetric model for the disorder, but as it turns out the amount of algebra is somewhat reduced by the asymmetric definition, while switching from one form to the other does not present any difficulty. The propagator of the disordered medium $G$ relates to the one of a pure host medium (pure $B$-phase) $g$ through the Dyson equation \begin{equation} G=g+gUG=g+gUg+gUGUg \label{a1}. \end{equation} The CPA for a disordered medium is introduced by the usual method \cite{bi1,bi2} of placing the impurities in a self-consistent medium such that the the averaged propagator of this effective medium fulfills the relation \begin{equation} \bar G=g+g \Sigma \bar G \label{a2} \end{equation} where $\Sigma$ is the CPA single particle self energy. $\Sigma$ itself is determined by the self consistency condition that the average of the total scattering of a particle in the effective medium be zero. This total scattering is described by the equation \begin{equation} G=\bar G + \bar G T \bar G \label{a8}. \end{equation} where $T$ is corresponding T-matrix of the problem and the self consistency condition is therefore $ \langle T \rangle \equiv 0$. In the SSA an additional requirement is made through the condition that the total scattering off a single site $n$ be zero. This scattering is described by the single site contribution to the T-matrix $T_n$ which is defined as \begin{equation} T_n = (U_n - \Sigma_n) \left[ 1 + F T_n \right] \label{a3} \end{equation} where $F$ is the site diagonal average propagator and $U_n$ and $\Sigma_n$ are the single site decompositions of $U$ and $\Sigma$, such that $U= \sum_n U_n$ and $\Sigma_n = \sum_n \Sigma_n$. The average of the disorder potential $U$ alone amounts to $\langle U \rangle = c V$ This along with the average of (\ref{a3}) set to zero determines $\Sigma$ to be \begin{equation} \Sigma = \frac {cV} {1-(V- \Sigma )F} \label{a3a}. \end{equation} One can now define conditional or weighted propagators \cite{bi26} which explicitly describe the propagation of a particle between partly or completely specified types of sites by multiplying a normalized version of the random potential onto them. For example, \begin{equation} G^i = \frac {U G} V \label{a4} \end{equation} describes the motion of a particle commencing on an impurity site and ending at an arbitrary other site in the medium, since $U$ will be zero if the first site of the function is a host. Similarly, \begin{equation} G^{ii}= \frac {UGU} {V^{2}} \label{a6} \end{equation} describes a situation where both sites are required to be impurities for the function not to be zero. Upon averaging over all configurations, the Green's functions become translationally invariant and the following relationships between the averaged weighted and unweighted functions can be obtained \begin{eqnarray} \bar G^i & = & \frac \Sigma V \bar G \label{a5} \\ \bar G^{ii} & = & \left( \frac \Sigma V \right)^2 \bar G + \frac {\Sigma - cV} {V^2}. \label{a7} \end{eqnarray} The second term corrects the site diagonal elements when $\bar G^{ii} = \bar G^i$ and uses identity (\ref{a3a}). From here on, other weighted functions can be calculated by probability conservation. It is found that \begin{eqnarray} \bar G^h & = & \bar G - \bar G^i = \left( 1- \frac \Sigma V \right) \bar G \label{a5a} \\ \bar G^{ih} & = & \bar G^{hi} = \bar G^i - \bar G^{ii} = \frac \Sigma V \left( 1 - \frac \Sigma V \right) \bar G - \frac {\Sigma - cV} {V^2} \label{a7a} \\ \bar G^{hh} & = & \bar G^h - \bar G^{ih} = \left( \frac \Sigma V - 1\right)^2 \bar G + \frac {\Sigma - cV} {V^2}. \label{a8a} \end{eqnarray} It is our goal in the present paper to establish the two particle analogues of these weighted Green's functions, i.e. jointly averaged products of such functions including the coherent scattering which induces correlations in the joint propagation of two particles. In order to deduce these jointly averaged weighted functions, we first need to obtain the non-averaged weighted functions, starting with $G^i$ and $G^{ii}$, in a representation such that no products between the disorder potential $U$ and the unweighted single particle function $G$ occur. To accomplish this, one can simply employ equation (\ref{a1}) which yields \begin{eqnarray} G^i & = & \frac {g^{-1} G - 1} V \label{a9} \\ G^{ii} & = & \frac 1 {V^2} \left[ g^{-1}Gg^{-1} - g^{-1} - U \right] .\label{a10} \\ \end{eqnarray} and by means of (\ref{a8}) these go over to \begin{eqnarray} G^i & = & \frac 1 V \left[ g^{-1} \bar G + g^{-1} \bar G T \bar G -1 \right] \label{a11} \\ G^{ii}& = & \frac 1 {V^2} \left[g^{-1} \bar G g^{-1} + g^{-1} \bar G T \bar G g^{-1} - g^{-1} - U \right] \label{a12}. \end{eqnarray} A further single particle identity resulting from (\ref{a2}) which will prove to be very useful is is \begin{equation} g^{-1} \bar G = \bar G g^{-1} = \Sigma \bar G + 1 . \label{a13} \end{equation} \section{Two particle theory} In this section a general weighted two particle theory involving two different bands is established from which the case of two particles moving in a single band can also be immediately obtained. The calculation for the unweighted two particle function has been made in {\bf I}. As a main result of that work the appropriate vertex corrections for the CPA were obtained which account for the coherent scattering processes of two particles that arise in the otherwise non-interacting two particle function through the averaging process. We follow the outline of Velick{\'y}'s reasoning to obtain the proper weights for two choices of the functions $ \langle G^{ \mu \nu}_aCG^{\mu^{\prime} \nu^{\prime}}_b \rangle $. The labels $ \mu, \nu, \mu^{\prime}, \nu^{\prime} \in \{i,h,0\} $ indicate the kind of weight -- either an impurity, a host or no weight -- which is attributed to the first and second site of the respective function. $C$ represents a generalized operator coupling the two single particle functions and $a$ and $b$ label two possibly different bands on which the respective single particle resolvents are defined. The positions of the weights within one single Green's function are thereby important, since the disordered medium before averaging is neither homogeneous nor isotropic and thus the non-averaged Green's functions depend non-trivially on both arguments. In principle there now would be 80 (!) possible ways of applying specific weights to these functions before averaging. However, in most cases, even if different bands are involved, only two particle functions having an equal kind of weighting on its single particle constituents will be needed in most applications. Later, we will consider two particular examples of two particle functions which find frequent use in linear response theory where the operator $C$ is diagonal in real space and also only diagonal elements (or sums of diagonal elements) of the functions defined above are used. For both cases considered, the total choices of weightings reduce to only five different ones, since the first and the second site of the first function will be the same as the second and first site of the second function, respectively, which implies that they must also pairwise bear the same weighting label. As an example we calculate for a most general choice of $C$ only the two functions $ \langle G_a^{ii} C G_b^{ii} \rangle $ and $ \langle G_a^{i0} C G_b^{0i} \rangle $ in this section, since they will prove to be the most useful types of weightings for the cases discussed thereafter and all other ones could be obtained in complete analogy to the calculations presented here. Using the identities (\ref{a11}) and (\ref{a12}), the single weighted two particle function can be written in terms of the single particle functions as \begin{equation} \langle G^{i0}_a C G^{0i}_b \rangle = V_a^{-1} V_b^{-1} \langle \left[ g^{-1}_a \bar G_a + g^{-1}_a \bar G_a T_a \bar G_a -1 \right] C \left[ \bar G_b T_b \bar G_b g^{-1}_b + \bar G_b g_b^{-1} -1 \right] \rangle \label{b4} \end{equation} and the double weighted one as \begin{eqnarray} \langle G^{ii}_aCG^{ii}_b \rangle & = & V_a^{-2} V_b^{-2} \langle \left[(g_a^{-1} \bar G_a g_a^{-1} -g_a^{-1}) + g_a^{-1} \bar G_a T_a \bar G_a g_a^{-1} - U_a \right]C \nonumber \\ & \times & \left[(g_b^{-1} \bar G_b g_b^{-1} - g_b^{-1}) + g_b^{-1} \bar G_b T_b \bar G_b g_b^{-1} - U_b \right] \rangle . \label{b5} \end{eqnarray} We calculate the double weighted function first, since it provides the more difficult task and from its solution it is straightforward to derive the one for the single weighted function as well. The problem of evaluating (\ref{b5}) is divided into two parts, the first one involving all terms not containing the matrix $U$, and the second one containing the remainder, i.e. \begin{equation} V_a^2V_b^2\langle G^{ii}_aCG^{ii}_b \rangle = {\cal K} + {\cal M} \label{b6} \end{equation} \begin{eqnarray} {\cal K} & = & \left[g_a^{-1} \bar G_a g_a^{-1} - g_a^{-1}\right] C \left[g_b^{-1} \bar G_b g_b^{-1} -g_b^{-1}\right] + g_a^{-1} \bar G_a \langle T_a \bar G_a g_a^{-1}C g_b^{-1} \bar G_b T_b \rangle \bar G_b g_b^{-1} \label{b7} \\ {\cal M} & = & \langle U_aCU_b \rangle + \left[ g^{-1}_a C \langle U_b \rangle - g^{-1}_a \bar G_a g^{-1}_a C \langle U_b \rangle - g^{-1}_a \bar G_a \langle T_a \bar G_a g^{-1}_a C U_b \rangle + (a \leftrightarrow b) \right] \label{b8} \end{eqnarray} where in ${\cal K}$ the terms involving an average over a single T-matrix have vanished, which is the standard CPA condition and $(a \leftrightarrow b)$ indicates that the labels are exchanged and the corresponding expressions reflected around the operator $C$. We evaluate ${\cal K}$ first since it is the term needed in the wider range of applications. With the identity (\ref{a13}) we find that \begin{equation} \left[g^{-1}_a \bar G_a g^{-1}_a - g^{-1}_a\right] C \left[g^{-1}_b \bar G_b g^{-1}_b -g^{-1}_b\right] = \Sigma_a \Sigma_b g_a^{-1} \bar G_a C \bar G_b g^{-1}_b \label{b9}. \end{equation} The T-matrix can be decomposed into its single site contributions $T_n$ as \begin{equation} T = \sum_n T_n + \sum_{n \not= m} T_n \bar G T_m + \sum_{n \not= m \not= l} T_n \bar G T_m \bar G T_l + ... \label{b11} \end{equation} Thereby the characteristic exclusions in the sums prevent the particle from scattering twice in sequence on the same site and $T_n$ satisfies equation (\ref{a3}). As shown by Velick{\'y}, $T$ can then be replaced in two ways by a closed set of equations, namely \begin{equation} T = \sum_n Q_n = \sum_n \tilde Q_n \label{b12} \end{equation} where \begin{equation} Q_n= T_n(1+ \bar G \sum_{n \not= m} Q_m) \label{b13} \end{equation} and \begin{equation} \tilde Q_n = (1+ \sum_{n \not= m} \tilde Q_m \bar G) T_n. \label{b13_5} \end{equation} Due to the requirement that $\langle T \rangle =0$ and the single site decomposition of $T$ from (\ref{b11}) it is possible to decompose averages on different sites to give \begin{equation} 0 \equiv \langle T \rangle = \sum_n \langle Q_n \rangle = \sum_n \langle T_n \rangle \left( 1 + \bar G \sum_{m \not= n} \langle Q_m \rangle \right) \label{b14} \end{equation} This also implies that $\langle Q_n \rangle = \langle \tilde Q_n \rangle = 0$. A vertex function $\Gamma$ can now be defined similar to {\bf I} such that \begin{equation} {\cal K} = g^{-1}_a \bar G_a ( \Sigma_a \Sigma_b C + \Gamma ) \bar G_b g^{-1}_b \label{b14_5} \end{equation} where now \begin{equation} \Gamma = \langle T_a \bar G_a g^{-1}_a C g^{-1}_b \bar G_b T_b \rangle . \label{b15} \end{equation} $\Gamma$ can be manipulated along the lines of {\bf I} by using (\ref{b12}) to yield \begin{equation} \Gamma=\sum_n \sum_m \langle Q_n^a \bar G_a g^{-1}_a C g_b^{-1} \bar G_b \tilde Q_m^b \rangle \label{b16}. \end{equation} By means of (\ref{b13}) - (\ref{b14}) this can be cast into the form \begin{equation} \Gamma_n = \langle T_n^a \bar G_a g^{-1}_a(C + g_a \sum_{p \not= n} \Gamma_p g_b)g^{-1}_b \bar G_b T_n^b \rangle \label{b17} \end{equation} where now $\Gamma = \sum_n \Gamma_n$, since from the decoupling introduced in (\ref{b14}) one gets $ \langle Q_n^a \bar G_a g^{-1}_a C g_b^{-1} \bar G_b \tilde Q_m^b \rangle = \langle Q_n^a \bar G_a g^{-1}_a C g_b^{-1} \bar G_b \tilde Q_n^b \rangle \delta_{n,m}$. The only difference to the corresponding expression in {\bf I}, (cf. (47) there) is that $\sum_{p \not= n} \Gamma_p$ is surrounded by the propagators of the pure medium $g_{a/b}$ here. We can then use (\ref{b14_5}) and (\ref{b17}) to obtain \begin{equation} \Gamma_n = \langle T_n^a \bar G_a g^{-1}_a C g^{-1}_b \bar G_b T_n^b \rangle + \langle T_n^a g_a{\cal K}g_b T_n^b \rangle - \Sigma_a \Sigma_b \langle T_n^a \bar G_a C \bar G_b T_n^b \rangle - \langle T_n^a \bar G_a \Gamma_n \bar G_b T_n ^b \rangle \label{b19}. \end{equation} At this stage we are able to find $\Gamma_n$ and therefore also ${\cal K}$. As we cast our model into a site representation we obtain $T_n = \mid n \rangle t_n \langle n \mid$ and hence \begin{equation} \Gamma_n=\mid n \rangle \Lambda \langle n \mid \left[ g_a{\cal K}g_b + \bar G_a g^{-1}_a C g^{-1}_b \bar G_b - \Sigma_a \Sigma_b \bar G_a C \bar G_b \right] \mid n \rangle \langle n \mid \label{b20} \end{equation} where $\Lambda$ is the irreducible vertex part derived by Velick{\'y} \begin{equation} \Lambda(z_1,z_2) = \frac {\langle t_n^a(z_1) t_n^b(z_2) \rangle} {1+ F_a(z_1) \langle t_n^a(z_1) t_n^b(z_2) \rangle F_b(z_2) } \label{b21}. \end{equation} The vertex $\Lambda$ can be regarded as being intrinsic to the CPA, since it does not depend on the particular form of the operator $C$. Substituting (\ref{b20}) into (\ref{b14_5}) yields \begin{eqnarray} & &{\cal K}= \Sigma_a \Sigma_b g^{-1}_a \bar G_a C \bar G_b g^{-1}_b + \nonumber \\ & + & \Lambda g^{-1}_a \bar G_a \sum_n \mid n \rangle \langle n \mid \left[ g_a{\cal K}g_b + \bar G_a g^{-1}_a C g^{-1}_b \bar G_b - \Sigma_a \Sigma_b \bar G_a C \bar G_b \right] \mid n \rangle \langle n \mid \bar G_b g^{-1}_b. \label{b22} \end{eqnarray} Multiplying by $g_a$ and $g_b$ from the left and right, we solve for the diagonal elements $ \langle m \mid g_a{\cal K}g_b \mid m \rangle $ \[ \langle m \mid g_a{\cal K}g_b \mid m \rangle = \Sigma_a \Sigma_b \langle m \mid \bar G_a C \bar G_b \mid m \rangle + \] \begin{equation} + \Lambda \sum_n \langle m \mid \bar G_a \mid n \rangle \langle n \mid \left[ g_a{\cal K}g_b + \bar G_a g^{-1}_a C g^{-1}_b \bar G_b - \Sigma_a \Sigma_b \bar G_a C \bar G_b \right] \mid n \rangle \langle n \mid \bar G_b \mid m \rangle \label{b23} \end{equation} and hence also solve (\ref{b22}). At this point it is helpful to visualize the form of the operator $C$ in the site representation, which in the most general case can be written as \begin{equation} C = \sum_{l,m} \gamma_{lm} \mid l \rangle \langle m \mid \label{b24} \end{equation} For convenience we introduce the short notation \begin{eqnarray} F_{n-m} & = & \langle n \mid \bar G \mid m \rangle ,\label{b25} \\ {\cal F}_{n-m} & = & \langle n \mid g^{-1} \bar G \mid m \rangle \label{b26} \end{eqnarray} where $F_0 \equiv F$ as already defined in (\ref{a3a}). It should be noted that $F_{n-m}$ and ${\cal F}_{n-m}$ are of different dimensions and their definition has been chosen to reduce the algebra as much as possible. Furthermore, these expressions show that equation (\ref{b23}) only contains translationally invariant quantities and hence it can be solved by Fourier transformation. The following Fourier transforms are introduced \begin{eqnarray} a_k & = & \Sigma_a \Sigma_b \sum_m e^{-ikR_m} \langle m \mid \bar G_a C \bar G_b \mid m \rangle \label{b28} \\ \alpha_k & = & \sum_m e^{-ikR_m} \langle m \mid g_a^{-1} \bar G_a C \bar G_b g^{-1}_b \mid m \rangle = \nonumber \\ & = & \sum_m e^{-ikR_m} \left[ \Sigma_a \Sigma_b \langle m \mid \bar G_a C \bar G_b \mid m \rangle + \gamma_{mm} + \Sigma_a \sum_n \gamma_{nm} F^a_{m-n} + \Sigma_b \sum_n \gamma_{mn} F^b_{n-m} \right] \label{b32} \\ A_k & = & \sum_m e^{-ikR_m} F^a_m F^b_{-m} \label{b30} \\ {\cal A}_k & = & \sum_m e^{-ikR_m} {\cal F}^a_m {\cal F}^b_{-m} \label{b31} \\ b_k & = & \sum_m e^{-ikR_m} \langle m \mid g_a{\cal K}g_b \mid m \rangle \label{b33} \\ c_k & = & \sum_m e^{-ikR_m} \langle m \mid {\cal K} \mid m \rangle \label{b34}. \end{eqnarray} The units of $a_k, \alpha_k, b_k, {\cal A}_k$ are unity, that of $A_k$ is $J^{-2}$ and that of $c_k$ is $J^2$. Inserting into (\ref{b23}) yields \begin{equation} b_k = \frac {a_k + \Lambda A_k (\alpha_k - a_k)} {1 - \Lambda A_k} \label{b34_5}, \end{equation} and hence by means of (\ref{b22}) \begin{equation} c_k = \alpha_k \frac {\Sigma_a \Sigma_b + \Lambda ( {\cal A}_k - \Sigma_a \Sigma_b A_k ) } {1 - \Lambda A_k} = \alpha_k \frac {\Sigma_a \Sigma_b + \Lambda D_k } {1 - \Lambda A_k} \label{b35} \end{equation} where \begin{equation} D_k = {\cal A}_k - \Sigma_a \Sigma_b A_k = \sum_m e^{-ikR_m} \left[ \gamma_{mm} + \Sigma_a \sum_n \gamma_{nm} F^a_{m-n} + \Sigma_b \sum_n \gamma_{mn} F^b_{n-m} \right] \label{b37}. \end{equation} Similarly, one can evaluate the term ${\cal M}$ from (\ref{b8}), but it turns out that little simplification can be made until an explicit form of $C$ is known and therefore we postpone its evaluation to the next section. For now, we proceed to calculate the single weighted function $\langle G^{i0}_a C G^{0i}_b \rangle$. From (\ref{b4}) we get \begin{equation} V_a V_b \langle G^{i0}_a C G^{0i}_b \rangle = g^{-1}_a \bar G_a C \bar G_b g^{-1}_b + C - g^{-1}_a \bar G_a C - C \bar G_b g^{-1}_b + g^{-1}_a \bar G_a \langle T_a \bar G_a C \bar G_b T_b \rangle \bar G_b g^{-1}_b. \label{bb1} \end{equation} Using (\ref{a13}), this can be recast into \begin{equation} V_a V_b \langle G^{i0}_a C G^{0i}_b \rangle = \Sigma_a \Sigma_b \bar G_a C \bar G_b + g^{-1}_a \bar G_a \langle T_a \bar G_a C \bar G_b T_b \rangle \bar G_b g^{-1}_b \label{bb2} \end{equation} which can be readily solved since the term $\langle T_a \bar G_a C \bar G_b T_b \rangle$ exactly corresponds to the vertex part of the unweighted function which is known from {\bf I} \begin{equation} \langle T_a \bar G_a C \bar G_b T_b \rangle = \sum_n \mid n \rangle \Lambda \langle n \mid \langle G_a C G_b \rangle \mid n \rangle \langle n \mid \label{bb2.5}. \end{equation} Thus, the Fourier transform \begin{equation} d_k = V_a V_b \sum_m e^{ikR_m} \langle m \mid \langle G^{i0}_a C G^{0i}_b \rangle \mid m \rangle \label{bb3} \end{equation} can be obtained using the solution for the unweighted two particle function and equations (\ref{b28}), (\ref{b30}) and (\ref{b31}) to yield \begin{equation} d_k = a_k + \Lambda {\cal A}_k \frac {a_k} { \Sigma_a \Sigma_b (1 - \Lambda A_k)} \label{bb4}. \end{equation} Using (\ref{b37}) this can be written as \begin{equation} d_k = \frac {a_k} {\Sigma_a \Sigma_b} \frac {\Sigma_a \Sigma_b + \Lambda D_k} {1 - \Lambda A_k }. \label{bb5} \end{equation} It is important to indicate at this point that the solution for a single weighted function in which the weights have been swapped to the inside, i.e. $ \langle G^{0i}_a C G^{i0}_b \rangle $, will not be the same as the one just obtained. A similar type of calculation yields instead of (\ref{bb4}) \begin{equation} d_k^{ \prime } = a_k + \Lambda A_k \frac {\alpha_k} {1 - \Lambda A_k}, \label{bb6} \end{equation} where $ d_k^{ \prime } = \sum_n e^{ikR_m} \langle m \mid \langle G^{0i}_a C G^{i0}_b \rangle \mid m \rangle $. Since we have now obtained expressions for the Fourier transformed site diagonal elements of some of the impurity weighted two particle functions, we are also able to find their respective off diagonal parts by inserting into the appropriate Bethe-Salpeter equations ((\ref{b22}) in the case of the double weighted function). Other types of weighted functions can be calculated from here by setting up the corresponding equations in analogy to (\ref{b4}) and (\ref{b5}) from the single particle functions. In particular, to obtain the corresponding host weighted functions one can apply the weights $(1 - U_{a/b}/V_{a/b})$ to the single particle resolvents from either side, but the form of this weight already suggests that host weighted properties can always be deduced by adding and/or subtracting corresponding unweighted/impurity weighted functions. For the most general case the set of closed equations will be quite large and for that reason we will in the following deduce these equations only in some more specific cases. \section{Applications in linear response theory} The most obvious application of a two particle theory as introduced above is in linear response theory, i.e. in calculating weighted susceptibilities. The study of weighted functions in this context helps to determine how such quantities as susceptibilities and transport coefficients are constituted (on the average) from processes on different components of the alloy. A further advantage of such a differentiation is that it allows for a more refined treatment of further renormalizations to the considered quantities once further interactions are introduced into the problem. The standard expression employed in linear response theory is a generalized Kubo formula \begin{equation} \chi_{C^{(1)},C^{(2)}}(z) = \int d \xi d \eta S_0 ( z,\xi,\eta ) Z^{-1} Tr_b \{ e^{- \beta H } C^{\dagger (2)} \langle \hat \delta^{\mu \nu} (\xi - H_a ) C^{(1)} \hat \delta^{ \mu \nu } (\eta - H_b ) \rangle \} . \label{c1} \end{equation} Thereby $ \chi_{C^{(1)},C^{(2)}} $ stands for a generalized type of susceptibility characterizing the linear response of an observable $C^{(1)}$ to an external perturbation coupling into the Hamiltonian through the operator $C^{(2)}$. $S_0$ denotes the zero order associated two particle Matsubara function in the Lehmann representation \cite{bi25} and $\hat \delta^{ \mu \nu} (\xi - H_{a/b}) $ is the corresponding spectral function of the full resolvent $ G^{ \mu \nu}_{a/b} ( \xi ) = \Xi^{\mu}_{a/b} [ \xi - H_{a/b} ]^{-1} \Xi^{\nu}_{a/b}$, defined on the subspace (band) $ a/b $, with the weighting operators $\Xi^{\mu}_{a/b}$ applied to it which assume the values $\Xi^i_{a/b} = U_{a/b}/V_{a/b}$, $\Xi^h_{a/b} = (1-U_{a/b}/V_{a/b})$ and $\Xi^0_{a/b} = 1$. $Tr_b$ denotes the trace over the subspace (band) $b$. Our particular interest focuses on the operators $C^{(1)}$ and $C^{(2)}$ which we now take to have the following diagonal form \begin{equation} C^{(1)/(2)} = \sum_m \gamma_m^{(1)/(2)} (\xi, \eta) \mid m \rangle \langle m \mid \label{c2a}. \end{equation} To examine the behavior of the weighted functions further under such a constraint it is useful to consider the two cases: \begin{equation} \gamma_m = const \times \delta_{m,r} \label{c2} \end{equation} where $r$ is an explicitly specified site and \begin{equation} \gamma_m = const . \label{c3} \end{equation} The first case arises in a treatment of the response from local interactions such as for example the space dependent exchange interaction between spin impurities (RKKY interaction \cite{bi21}) submerged in a system otherwise containing non-magnetic disorder \cite{bi22}. The second case is widely used in calculations of the optical response in metals and semiconductors\cite{bi15,bi20}, since the exciting optical fields can be taken as uniform over space with approximate momentum $q=0$ and the optical matrix elements describing transitions between bands of different angular momentum symmetry are usually well approximated as constants. In both cases the restriction to only consider diagonal elements of two particle functions reduces the set of closed equations for all possible weightings to \begin{equation} \langle G^h_a C G^h_b \rangle = \langle G_a C G_b \rangle - \langle G^i_a C G^i_b \rangle \label{c4} \end{equation} \begin{equation} \langle G^{ih}_a C G^{ih}_b \rangle = \langle G^{hi}_a C G^{hi}_b \rangle = \langle G^i_a C G^i _b\rangle - \langle G^{ii}_a C G^{ii}_b \rangle \label{c5} \end{equation} \begin{equation} \langle G^{hh}_a C G^{hh}_b \rangle = \langle G^h_a C G^h_b \rangle - \langle G^{ih}_a C G^{ih}_b \rangle \label{c6} \end{equation} It follows, that for this case the calculation of only $ \langle G^{ii}_aCG^{ii}_b \rangle $ and $ \langle G^i_aCG^i_b \rangle $ is sufficient to also obtain the remaining mixed and host weighted functions. We start to consider the case of (\ref{c2}) first. \subsection{Electronic Susceptibility} The response of a system at some point $r_1$ in space to a perturbation applied at point $r_2$ has a wide range of applicability. Disregarding the coupling constants which simply scale the result we assume that $C^{(1)} = \mid 2 \rangle \langle 2 \mid $ and $C^{(2)} = \mid 1 \rangle \langle 1 \mid $, where the choice of the origin is arbitrary. In a slightly different notation the calculation of the double weighted function corresponds to evaluating $\langle G_a^{ii}(r_1,r_2;t) G_b^{ii}(r_2,r_1;t^{ \prime }) \rangle$. For the RKKY interaction mentioned before, which couples spins at a distance $r_1 - r_2$ through the electronic spin susceptibility of electrons in the conduction band, the two-particle time dependent function is to be Fourier transformed and to be taken at two identical single particle energies \cite{bi21}. If the spins are at impurity sites, the weighted functions must be used to describe the problem adequately. Earlier treatments \cite{bi22} neglected the vertex corrections and it is proposed to investigate the effect of their inclusion in a separate publication. In this as in other problems the self-interaction $r_1=r_2$ can be neglected for most purposes, which implies that it will be sufficient to calculate only the term ${\cal K}$ from (\ref{b7}), since all terms occurring in ${\cal M}$ from (\ref{b8}) will vanish for $r_1 \not= r_2$, since the disorder potential $U$ is a diagonal matrix which vanishes identically for off-diagonal terms. To clarify the meaning of the quantity which is obtained through this special choice of $C^{(1)}$ and $C^{(2)}$, we look at how the corresponding Bethe-Salpeter (B-S) equation can be rewritten for the case of a unweighted two particle Green's function as introduced in equations (22) of Ref. \onlinecite{bi8}. In the single site approximation the B-S equation for the unweighted function reads \begin{eqnarray} \langle G_a(1,2) G_b(1,2) \rangle & \equiv & \langle G_{ab}^{(2)}(1,1;2,2) \rangle = \langle G_a(1,2) \rangle \langle G_b(1,2) \rangle + \nonumber \\ & + & \Lambda \sum_n \langle G_a(1,n) \rangle \langle G_b(1,n) \rangle \langle G_{ab}^{(2)}(n,n;2,2) \rangle \label{ca1} \end{eqnarray} where we have rewritten \begin{equation} \langle 2 \mid \bar G_a \mid 1 \rangle \langle 1 \mid \bar G_b \mid 2 \rangle \equiv \bar G_a(1,2) \bar G_b(1,2) \label{ca2} \end{equation} and \begin{equation} \langle 2 \mid \langle G_a \mid 1 \rangle \langle 1 \mid G_b \rangle \mid 2 \rangle \equiv \langle G_{ab}^{(2)}(1,1;2,2) \rangle \label{ca3}. \end{equation} For the Fourier transforms from before (\ref{b28}) - (\ref{b31}) this amounts to \begin{eqnarray} a_k & = & \Sigma_a \Sigma_b \sum_m e^{-ikR_m} \bar G_a(m) \bar G_b(m) \label{ca4} \\ \frac {a_k} {\Sigma_a \Sigma_b} & = & A_k \label{ca5} \\ \alpha_k & = & \sum_m e^{-ikR_m} \langle m \mid g_a^{-1} \bar G_a \mid 1 \rangle \langle 1 \mid \bar G_b g_b^{-1} \mid m \rangle = \nonumber \\ & = & \sum_m e^{-ikR_m} \left[ \Sigma_a \Sigma_b \bar G_a(m) \bar G_b(m) + \delta_{m,1} (1 + \Sigma_a F_a + \Sigma_b F_b) \right] = {\cal A}_k .\label{ca6} \end{eqnarray} Using (\ref{ca5}) and (\ref{ca6}) in connection with (\ref{bb4}) and (\ref{bb6}) the result for the single weighted functions is independent of whether the impurity weights are applied to the interior or to the exterior of the single particle resolvents, such that $ \langle G_a^{i0}(1,2) G_b^{0i}(1,2)\rangle = \langle G_a^{0i}(1,2) G_b^{i0}(1,2) \rangle $. Indeed it is not hard to show that other arbitrary distributions of two impurity weights also give the same results as long as there is one weight applied to each resolvent. The same is of course true for host related properties. Accordingly, $b_k$ from (\ref{b34_5}) in the previous section goes over to \begin{equation} b_k = a_k \frac {\Sigma_a \Sigma_b + \Lambda (\alpha_k - a_k)} {\Sigma_a \Sigma_b - \Lambda a_k} \label{ca7} \end{equation} and similarly $c_k$ from equation (\ref{b35}) to \begin{equation} c_k = \Sigma_a \Sigma_b \alpha_k \left( 1+\Lambda \frac {\alpha_k} {\Sigma_a \Sigma_b - \Lambda a_k} \right) \label{ca8}. \end{equation} We remember that the Fourier transform of the unweighted two particle Greens function is \begin{equation} \sum_{R_{1,2}} e^{ikR_{1,2}} \langle G_a(1,2) G_b(1,2) \rangle = \frac {a_k} {\Sigma_a \Sigma_b - \Lambda a_k } . \label{ca9} \end{equation} Phenomenologically, one can scrutinize the uncorrelated limit where $ \Lambda \rightarrow 0$ for which in (\ref{ca8}) the $k$- dependent part of $c_k$ behaves as \begin{equation} \lim_{\Lambda \rightarrow 0} c_k = \Sigma_a \Sigma_b a_k + ... \label{ca10} \end{equation} which is the correct limiting result for a product of two averaged double impurity weighted single particle functions $ \bar G_a^{ii}(1,2) \bar G_b^{ii}(1,2)$ or the weighted two particle function without coherent corrections. However, taking this limit is reasonable only in special cases since generally the vertex $\Lambda$ depends on the self energy $\Sigma$. It will be appropriate to take in a weak disorder limit (virtual crystal limit) since for $V_{a/b} \rightarrow 0$ the CPA predicts that $\Sigma_{a/b} \rightarrow c V_{a/b}$ and $\Lambda \rightarrow c (1-c) V_a V_b$ which means that $\Lambda$ approaches zero faster than $\Sigma_{a/b}$ in this limit. One should also note at this point, that although it is apparent that the term $(1+ \Sigma_a F_a + \Sigma_b F_b)$ , which arises in the difference of the Fourier transforms $\alpha_k$ and $a_k$, originates from the difference in the diagonal parts of their respective real space Green's functions, it can not be neglected in this treatment. When $\Lambda$ is finite, this term is multiplied with other $k-$dependent quantities and thus contributes to the off diagonal elements as well. Written in terms of $a_k$, (\ref{ca8}) can be cast into \begin{equation} c_k = \frac {a_k} { \Sigma_a \Sigma_b - \Lambda a_k} \left( \Sigma_a \Sigma_b + \Sigma_a \Sigma_b (a_k)^{-1} (1 + \Sigma_a F_a + \Sigma_b F_b) \right) \left( \Sigma_a \Sigma_b + \Lambda (1 + \Sigma_a F_a + \Sigma_a F_b) \right) \label{ca11}. \end{equation} If again diagonal terms in real space are neglected, i.e. $k-$independent quantities in $k-$space, it can be shown that (\ref{ca11}) can be cast into the compact form \begin{equation} \zeta_k = \frac {a_k} {\Sigma_a \Sigma_b - \Lambda a_k} {\left( \Sigma_a \Sigma_b + \Lambda (1 + \Sigma_a F_a + \Sigma_b F_b) \right)}^2 \label{ca12} \end{equation} where we have introduced $\zeta_k \equiv c_k + c_o$, whereby $c_o$ is independent of $k$. According to the definition of $\zeta_k$, this finally relates the weighted off-diagonal real space two particle Green's function to the unweighted one as \begin{equation} \langle G_a^{ii}(1,2) G_b^{ii}(1,2) \rangle = \langle G_a(1,2) G_b(1,2) \rangle \frac {{\left( \Sigma_a \Sigma_b + \Lambda (1 + \Sigma_a F_a + \Sigma_b F_b) \right)}^2} {(V_a V_b)^2} . \label{ca13} \end{equation} Equally, the relation for the single weighted function $\langle G_a^i(1,2) G_b^i(1,2)\rangle$ can be obtained almost immediately if (\ref{bb4}) is modified for this choice of $C^{(1)}$ and $C^{(2)}$, which goes over to \begin{equation} d_k = a_k + \Lambda \alpha_k \frac {a_k} { \Sigma_a \Sigma_b - \Lambda a_k} = \frac {a_k} { \Sigma_a \Sigma_b - \Lambda a_k} \left[ \Sigma_a \Sigma_b + \Lambda (1 + \Sigma_a F_a + \Sigma_b F_b ) \right] \label{ca14} \end{equation} thus relating the real space functions in this case as \begin{equation} \langle G_a^i(1,2) G_b^i(1,2) \rangle = \langle G_a(1,2) G_b(1,2) \rangle \frac { \Sigma_a \Sigma_b + \Lambda (1 + \Sigma_a F_a \Sigma_b F_b )} {V_a V_b} \label{ca15} \end{equation} where no diagonal contributions were omitted. The fact that in the single weighted case the same weighting factor occurs once as opposed to twice for the double weighted one is structurally equivalent to the results for the single particle theory. Although retrospectively that might not be surprising, it is also quite interesting in view of the fact that for the most general case of the previous section the weighting factors in (\ref{b35}) and (\ref{bb5}) almost look alike were it not for the difference in the pre-factors, i.e. $\alpha_k$ in the former and $a_k / \Sigma_a \Sigma_b$ in the latter case. In principle all other functions can be derived now using the set of relations (\ref{c4}) - (\ref{c6}). However, one can save a considerable amount of algebra by recalling the following relation which holds in the transition from impurity to host related properties in the CPA \begin{equation} - \Sigma \rightarrow (V - \Sigma) \hspace{20mm} c \rightarrow (1-c). \label{ca16} \end{equation} Starting from the definition of the vertex $ \Lambda $ in (\ref{b21}) and the CPA condition $ \langle T_n \rangle = 0 $ one finds that \begin{equation} \Lambda = \left[ \frac {1-c} {(V_a - \Sigma_a)(V_b - \Sigma_b)} + \frac c { \Sigma_a \Sigma_b } \right]^{-1} \label{ca17} \end{equation} which in a single band case ($a=b$) readily simplifies to the form first introduced by Leath \cite{bi8} \begin{equation} \Lambda_{a=b} = \frac {\delta \Sigma} {\delta F} = \frac { \Sigma ( V - \Sigma ) } { 1- ( V - 2 \Sigma) F }. \label{ca18} \end{equation} With the relation (\ref{a3a}) between $F$ and $\Sigma$ the weighting factor for the single weighted impurity function from (\ref{ca15}) reduces to \begin{equation} \frac { \Sigma_a \Sigma_b + \Lambda (1 + \Sigma_a F_a + \Sigma_b F_b) } {V_a V_b} = \frac {(1-c) \Lambda} {(V_a - \Sigma_a)(V_b - \Sigma_b)}. \label{ca19} \end{equation} Correspondingly, the factor for the double weighted function is the square of this quantity. One can now employ (\ref{c4}) and (\ref{c5}) to find the weights for the single host- and the impurity-host weighted functions, respectively. We find \begin{equation} \langle G_a^h G_b^h \rangle = \langle G_a G_b \rangle \frac {c \Lambda^2} {\Sigma_a \Sigma_b} \label{ca20} \end{equation} and \begin{equation} \langle G_a^{ih} G_b^{ih} \rangle = \langle G_a G_b \rangle \frac {c (1-c)\Lambda} {\Sigma_a \Sigma_b (V_a - \Sigma_a ) (V_b - \Sigma_b )} \label{ca21} \end{equation} which by means of (\ref{c6}) gives \begin{equation} \langle G_a^{hh} G_b^{hh}\rangle = \langle G_a G_b \rangle \left( \frac {c \Lambda} { \Sigma_a \Sigma_b } \right)^2. \label{ca22} \end{equation} Note that this could also have been expected from the transformation property (\ref{ca16}). The impurity and host weights as represented in (\ref{ca19}) and (\ref{ca20}) are immediately seen to be the two contributions summing to $ \Lambda^{-1} $ in (\ref{ca17}). Thus, even though there are many different possible representations of the two band vertex, the representation in (\ref{ca17}) shows that the transition from impurity to host properties leaves $\Lambda$ invariant. A notable feature about the weights calculated in this section is that they are independent of any wavevectors and only multiply the unweighted function as scalar energy dependent factors. This is a direct consequence of the single site approximation. \subsection{Theory of Absorption} In this section we evaluate a form of the two particle functions needed in the calculation of the linear response absorption in a disordered solid. We take the operators $C^{(1)}$ and $C^{(2)}$ in the characteristic form of dipole operators similar to the one in (\ref{c3}). Furthermore, we assume that the dipole matrix elements be essentially constant, such that $ C^{(1)/(2)}= \gamma^{(1)/(2)} \sum_m \mid m \rangle \langle m \mid $. This choice corresponds to applications for the description of processes involving transitions between bands of different angular momentum symmetry such as required by the selection rule for optical processes at at zero total momentum \cite{bi8,bi10}. At the end of this sub-section we will also give for completeness a short account of the calculation of linear response conductivities in disordered solids. For the above choice of $\gamma^{(1)/(2)}$ the calculation of the term ${\cal K}$ in (\ref{b22}) is greatly simplified compared to before, but it will be necessary now to also consider total diagonal terms, since the sums over all states in $C^{(1)/(2)}$ couple all sites and hence all contributions coming from term ${\cal M}$ in (\ref{b8}) have to be included. As a consequence of the introduction of the dipole operator the main change arising in the result for ${\cal K}$ is that the site diagonal elements $\langle n \mid {\cal K} \mid n \rangle $ and $\langle n \mid G_a C^{(1)} G_b \mid n \rangle$ as well as $\langle n \mid G_a g_a^{-1} C^{(1)} g_b^{-1} G_b \mid n \rangle $ are now actually independent of $n$ \cite{bi15}. Since $C^{(1)}$ now couples the functions to its left and right like a matrix product, the B-S equations (\ref{b22}) and (\ref{b23}) have a very simple solution in terms of their Fourier transforms. Introducing $a,b,c \equiv a_k, b_k, c_k \mid_{k=0}$ as the zero momentum elements of the respective transforms from last section we effectively get $c= \langle n \mid {\cal K}\mid n \rangle $ , $a = \Sigma_a \Sigma_b \langle n \mid \bar G_a \bar G_b \mid n \rangle $, where the omission of $C^{(1)}$ indicates that the two single particle resolvents are now simply multiplied as matrices, and equation (\ref{b34_5}) reduces to \begin{equation} b=a \frac {\Sigma_a \Sigma_b + \Lambda (\alpha - a)} {\Sigma_a \Sigma_b - \Lambda a} \label{cb1} \end{equation} and equation (\ref{b35}) to \begin{equation} c = \Sigma_a \Sigma_b \alpha \left( 1 + \Lambda \frac \alpha {\Sigma_a \Sigma_b - \Lambda a} \right) \label{cb2} \end{equation} which can be recast into \begin{eqnarray} c & = & \frac a {\Sigma_a \Sigma_b - \Lambda a} \left[ \Sigma_a \Sigma_b + \Lambda (1 + \Sigma_a F_a + \Sigma_b F_b) \right]^2 \nonumber \\ & + & (1 + \Sigma_a F_a + \Sigma_b F_b) \left[ \Sigma_a \Sigma_b + \Lambda (1 + \Sigma_a F_a + \Sigma_b F_b) \right] \label{cb3} \end{eqnarray} where the term independent of $a$ which was discarded in the previous section has to be kept in this case since the contributions of the diagonal elements become important. In calculating ${\cal M}$ one has to be aware that $U$ is a matrix which just has a random occupation of its diagonal. The sum over all sites in the operators $C^{(1)}$ and $C^{(2)}$ will hence just pick out the sum of all total diagonal parts $ \langle n \mid {\cal G} \mid n \rangle \langle n \mid U \mid n \rangle$ where ${\cal G}$ is a generalized product of several Green's functions in the same band (the case where $U$ and ${\cal G}$ are swapped is analogous). From (\ref{b8}) we get immediately \begin{equation} {\cal M} = c V_a V_b + \left[ cg_a^{-1}V_b - c g_a^{-1} \bar G_a g_a^{-1} V_b - g_a^{-1} \bar G_a \langle T_a \bar G_a g_a^{-1} U_b \rangle + (a \leftrightarrow b) \right] \label{cc1} \end{equation} where the only term giving slight complications is $\langle T_a \bar G_a g_a^{-1} U_b \rangle$. However by means of (\ref{a13}) we obtain \begin{equation} \langle T_a \bar G_a g_a^{-1} U_b \rangle = \Sigma_a \langle T_a \bar G_a U_b \rangle + \langle T_a U_b \rangle \label{cc2}. \end{equation} In this expression, the second term presents more complications, since for the first one we remember from (\ref{a8}) that \begin{equation} T \bar G = \bar G^{-1} G - 1 \label{cc3} \end{equation} and $U_b = U_a V_b/V_a$ such that \begin{equation} \langle T_a \bar G_a U_b \rangle = \langle \bar G_a^{-1} G_a U_b \rangle - c V_b = \frac { \Sigma_a V_b } {V_a} -cV_b. \label{cc4} \end{equation} Subsequently, the second term can be decoupled by means of (\ref{b12}) - (\ref{b13_5}) \begin{equation} \langle T_a U_b \rangle = \sum_n \langle \tilde Q_n^a U_b \rangle = \sum_n \langle (1 + \sum_{m \not= n} \tilde Q_m^a \bar G_a ) T_n^a U_b \rangle \label{cc5}, \end{equation} Applying (\ref{b14}) and using the the fact that $U$ is diagonal yields \begin{equation} \langle \tilde Q_m^a \bar G_a T_n^a U_b \rangle = \langle \tilde Q_m^a \rangle \bar G_a \langle T_n^a U_b \rangle = 0 \label{cc6} \end{equation} since $(m \not= n)$ and $\langle \tilde Q_m \rangle = 0$ and hence we find \begin{equation} \langle T_a U_b \rangle = \sum_n \langle T_n^a U_n^b \rangle = \frac {c (V_a- \Sigma_a)V_b} {1- (V_a- \Sigma_a) F_a}. \label{cc7} \end{equation} Collecting all terms for ${\cal M}$ and some more algebraic manipulation finally yields \begin{equation} {\cal M} = cV_a V_b - (1+ \Sigma_a F_a) \Sigma_a V_b - (1+ \Sigma_b F_b) \Sigma_b V_a. \label{cc8} \end{equation} The sum of all diagonal parts, i.e. ${\cal M}$ and the ones from the second term in (\ref{cb3}), can be shown to assume the very compact form \begin{equation} {\cal M} + (1 + \Sigma_a F_a + \Sigma_b F_b) \left[ \Sigma_a \Sigma_b + \Lambda (1 + \Sigma_a F_a + \Sigma_b F_b) \right] = \frac {\Lambda F_a F_b} { V_a V_b}. \label{cc9} \end{equation} The final result for $ \langle G_a^{ii} G_b^{ii} \rangle$ thus amounts to \begin{equation} \langle G_a^{ii} G_b^{ii} \rangle = \langle G_a G_b \rangle \left[ \frac {(1-c) \Lambda} {(V_a - \Sigma_a )( V_b - \Sigma_b)} \right]^2 + \frac {\Lambda F_a F_b} { V_a V_b}. \label{cc10} \end{equation} Here the first term has been rewritten in the same way as already derived in the last subsection for the finite range susceptibility. {}From there it is also seen that the single weighted function will have the same weight as calculated in (\ref{ca19}) for the corresponding function in the exchange coupling case. We find \begin{equation} \langle G_a^i G_b^i \rangle = \langle G_a G_b \rangle \frac {(1-c) \Lambda} {(V_a - \Sigma_a )( V_b - \Sigma_b)} \label{cc11} \end{equation} and by means of (\ref{c4}) - (\ref{c6}) \begin{equation} \langle G_a^h G_b^h \rangle = \langle G_a G_b \rangle \frac {c \Lambda} {\Sigma_a \Sigma_b} \label{cc12} \end{equation} \begin{equation} \langle G_a^{ih} G_b^{ih} \rangle = \langle G_a G_b \rangle \frac {c(1-c) \Lambda^2} {(V_a - \Sigma_a )( V_b - \Sigma_b) \Sigma_a \Sigma_b} - \frac {\Lambda F_a F_b} { V_a V_b} \label{cc13} \end{equation} \begin{equation} \langle G_a^{hh} G_b^{hh} \rangle = \langle G_a G_b \rangle \left[ \frac {c \Lambda} { \Sigma_a \Sigma_b} \right]^2 + \frac {\Lambda F_a F_b} { V_a V_b}. \label{cc14} \end{equation} Comparing the results of the last two sub-sections, it becomes clear that the several weights obtained are essentially universal. The main difference in the absorption case as compared to the susceptibility one comes from the diagonal terms which have to be kept in the double weighted functions. The single weighted analogues are void of this difficulty and the weighting factors are identical for both cases. So far we had omitted to consider a form of the two particle functions which is needed for conductivity calculations. However, Velick{\'y} \cite{bi4} showed for the unweighted functions that the vertex $\Gamma$ vanishes in the corresponding expression for the conductivity, due to the antisymmetry of the dipole matrix elements in $k$-space if they are taken between Bloch states of a non interacting Fermi system in a crystal with inversion symmetry. The same is also true for the weighted case and effectively the weighted functions which would have to be used for such calculations would just consist of products of the corresponding single particle quantities. This however turns out to be a general deficiency of the CPA in the single site approximation since due to the multiple scattering exclusions, only ladders of nested diagrams are used in calculating the total contribution of the coherent scattering. The CPA therefore neglects higher order two particle correlations which are in fact non-zero and contribute markedly to the conductivity. Langer and Neal \cite{bi11} have shown that the so called ``maximally crossed'' diagrams, i.e. diagrams which have a maximal crossing of coherent particle-particle scattering lines actually contribute the leading part -- in the order of the expansion considered -- to the full two particle disorder vertex for the conductivity in an otherwise non-interacting system. For the case of interacting Fermi systems, however, the presence of the interactions is sufficient to destroy the aforementioned antisymmetry and thus also the terms already included in the vertex of a single site two particle CPA as discussed here will give a finite contribution to the conductivity in real systems. \subsection{Split band limit} As already indicated, in contrast to many other theories of disorder, the CPA interpolates correctly to the limits of strong disorder and high concentrations. In this situation each band splits into two components of strength $c$ and $1-c$, respectively, which represent largely separate $A-$ and $B$-type excitations. An intuitive consideration of the underlying physics in this limit suggests, that the correct description of an absorption process should predict that the overlap integral for transitions between sites pertaining to different alloy components will gradually decrease and thus in reverse, that transitions between sites of equal type will be more and more favored. In the following, we prove that the CPA of weighted two particle functions predicts this behavior correctly which makes it useful for a better quantitative understanding of absorptive and dispersive processes in strongly disordered alloys. The corresponding single particle theory must fail in this respect, since it will weight the occurring transitions only with the products of concentrations of sites involved in these transitions. To illustrate this we assume that our material components $A$ and $B$ have corresponding single site energies $ \varepsilon_a^A $, $ \varepsilon_b^A $ and $ \varepsilon_a^B $, $ \varepsilon_b^B $ for the two bands respectively and the carriers have become totally localized, i.e. their effective mass goes to infinity, or vice versa the bandwidths involved go to zero. The potentials $V_a$ and $V_b$ are then defined as \begin{equation} V_a = \varepsilon^A_a - \varepsilon^B_a, \hspace{20mm} V_b= \varepsilon^A_b - \varepsilon^B_b \label{sb1} \end{equation} and the single particle site diagonal Green's functions go over to \begin{equation} F_{\lambda}(z) = \frac c {z-\varepsilon_{\lambda}^A} + \frac {1-c} {z-\varepsilon_{\lambda}^B} \equiv \frac 1 {z - \varepsilon_\lambda^B - \Sigma_{\lambda}} \label{sb2} \end{equation} where $\lambda$ labels the corresponding band. Thus, if a two particle theory is constructed from single particle properties only, and coherent terms in the two particle scattering are neglected, this leads to peaks in the absorption spectrum as shown in table \ref{Table1}. The energies at which the peaks are centered are shown in row 2 and their relative weight for the uncorrelated average is shown in the row 3. In the following we show for the limit of strong disorder how, upon introducing the vertex corrections in conjunction with the appropriate weighting factors for the two respective components, the expected transitions are filtered out correctly with their appropriate transition strengths and the spurious crossed terms are suppressed as shown in row 4 of table \ref{Table1}. Defining \begin{equation} z-\varepsilon_{\lambda}^B \equiv x \label{sb3} \end{equation} $F$ and $\Sigma$ can be rewritten as \begin{eqnarray} F(x) & = & \frac 1 {x- \Sigma} = \frac {x-(1-c)V} {x(x-V)} \label{sb4} \\ \Sigma(x)& = & \frac {cVx} {x-V(1-c)} \label{sb5} \end{eqnarray} and the vertex $\Lambda$ given by (\ref{ca17}) can be written \begin{equation} \Lambda = \frac {(1-c)c V_a V_b x_a x_b (x_a - V_a)(x_b-V_b)} { \left[ x_a - V_a(1-c) \right] \left[ x_b - V_b(1-c) \right] \left[ c x_a x_b + (1-c)(x_a-V_a)(x_b-V_b) \right]} \label{sb7}. \end{equation} The impurity weight $\xi \equiv (1 - c) \Lambda_{cv} / (V_c -\Sigma_c) (V_v - \Sigma_v)$ can be represented as \begin{equation} \xi = \frac {cx_ax_b} {cx_ax_b + (1-c)(x_a - V_a)(x_b - V_b)} \label{sb8} \end{equation} and equally the host weight $\eta \equiv c \Lambda_{ab} / \Sigma_a \Sigma_b$ as \begin{equation} \eta = \frac {(1-c)(x_a - V_a)(x_b - V_b)} {cx_ax_b + (1-c)(x_a - V_a)(x_b - V_b)}. \label{sb9} \end{equation} It is now evident, that the impurity weighted quantities are proportional to $c$ and the host weighted ones to $1-c$ and not the other way around as their the appearance in terms of their weighting factors might superficially have suggested. The correction factor coming from the diagonal terms in the double weighted functions can be recast into \begin{equation} \gamma \equiv \frac {\Lambda F_a F_b} {V_a V_b} = \frac {(1-c) c} { c x_a x_b + (1-c)(x_a-V_a)(x_b-V_b)} . \label{sb9.5} \end{equation} Furthermore, the unweighted but vertex corrected two particle propagator $K$ assumes the form \begin{equation} K= \frac { cx_ax_b + (1-c)(x_a - V_a)(x_b - V_b) } {x_ax_b (x_a - V_a)(x_b - V_b) }. \label{sb10} \end{equation} After further algebra one can show, that in this limit the weighted functions can be expressed as \begin{eqnarray} & & \langle G^{ii} G^{ii} \rangle = K \xi^2 + \gamma = \langle G^i G^i \rangle = K \xi = \frac c {(x_a-V_a)(x_b-V_b)} \label{sb11} \\ & & \langle G^{hh} G^{hh} \rangle = K \eta^2 + \gamma = \langle G^h G^h \rangle = K \eta = \frac {(1-c)} {x_ax_b} \label{sb12} \\ & & \langle G^{ih} G^{ih} \rangle = 0 \label{sb13_5} \end{eqnarray} which is exactly what is expected to happen physically in this limit. The crossed terms in the transition are canceled out -- hence the crossed function in (\ref{sb13_5}) goes to zero -- and the double weighted functions become identical to the single weighted ones, since now effectively only the totally site diagonal element $K_D \equiv \langle G^{ \nu \mu }(l,l) G^{ \nu \mu}(l,l) \rangle$ still contributes to the transitions, which implies that only two possibilities for weighting the two particle functions remain, namely as $\langle G^i_a G^i_b \rangle$ and $\langle G^h_a G^h_b \rangle$. As could be expected from a theory which properly describes the strong disorder limit the transition strengths now distribute with the concentrations $c$ and $1-c$ between the $A_b \rightarrow A_a$ and $B_b \rightarrow B_a$ transitions, respectively, such as shown in row 4 of table \ref{Table1}. This feature may in reverse be used to derive the total diagonal element $K_D$ for all ranges of disorder. By requiring $K^{ih}_D = 0$ we find \begin{equation} K_D = \frac \gamma {\xi\eta} = \frac {F_a F_b} {V_a V_b} \left( \frac { \Sigma_a \Sigma_b } c + \frac {(V_a - \Sigma_a ) (V_b - \Sigma_b)} {(1-c)} \right) \label{sb13}. \end{equation} The weighted versions of this element are obtained by just multiplying the corresponding single weights from (\ref{cc11}) and (\ref{cc12}) on to it. Moreover, $K_D$ is equivalent to the $r_{1,2}=0$ component of the two particle function calculated for the finite range susceptibility in the last sub-section. In terms of the notation introduced there it reads \begin{equation} K_D = \sum_k \frac {a_k} {\Sigma_a \Sigma_b - \Lambda_{a,b} a_k} \label{sb14} \end{equation} which would have been harder to evaluate starting from that representation. The total site diagonal element thus decouples into the corresponding site diagonal single particle functions with an appropriate correction term accounting for the coherent processes. \section{Numerical Results} In this subsection, in order to exemplify the general results, we discuss numerical results obtained for the optical absorption in a non-degenerate binary semiconductor alloy for a given model density of states. We are thus able to show how a CPA type of polarization, including vertex corrections, decomposes into contributions originating from single alloy components as the strength of the disorder is increased thus eventually causing the joint density of states to split into several components (up to three different ones for the double weighted case). In all our calculations we have used a semi elliptic density of states for a pair of 3-D conduction and valence bands as introduced in Ref. \onlinecite{bi3} for the single particle CPA, i.e. \begin{equation} \begin{tabular}{ c c } $ \rho_{\lambda} (E) = \displaystyle{ \frac 2 {\pi w_{\lambda}^2}} \sqrt{w_{\lambda}^2 - E^2} $ & $ \mid E \mid \leq w_{\lambda} $ \\ \\ $ \rho_{\lambda} (E) = 0 $ & $ \mid E \mid \leq w_{\lambda} $ \label{n1} \end{tabular} \end{equation} where $\lambda$ labels either the conduction $\lambda = a$ or valence $\lambda = b$, and $w_{\lambda}$ is the half-width of the band considered. This has the advantage that the self consistent CPA equation for the self energy $\Sigma (E)$ is a third degree polynomial which can be solved analytically. To understand the effects that arise from genuine two particle behavior as compared to those expected from the single particle CPA, we recapitulate some of the features of the single particle theory first, mainly building on the treatment presented in Ref. \onlinecite{bi3}. It is established there, that depending on the concentration and disorder strength relative to the bandwidth, an impurity band is eventually split off while in this split regime under some circumstances the CPA self energy exhibits a pole. Fig. \ref{FIG1} shows a reproduction of the ``phase'' diagram first presented there, indicating how the several regions are separated. It can be seen that for a disorder strength $ \mid V_{\lambda}/w_{\lambda} \mid > 1$ the bands always split into $A-$ and $B$-components, whereas the splitting occurs earlier as the concentration $c$ is reduced, going down to $\mid V_{\lambda}/w_{\lambda} \mid > 0.5$ as $c \rightarrow 0$. We have calculated the linear polarizability of the medium by employing a Kubo formula as introduced in (\ref{c1}). Furthermore, we continue assume that the optical matrix elements are essentially constant and that such elements are the same for both alloy components and we hence normalize them to unity. The optical absorption is the negative imaginary part of the retarded polarizability of the disordered medium $-\text{Im}\Pi(\omega ) $, which can be formally written as \begin{equation} \Pi_r( \omega ) = - \lim_{\stackrel {\scriptstyle \beta \rightarrow \infty}{i \omega \rightarrow \omega + i \delta }} \beta^{-1} \sum_{i \varepsilon} \int \frac {d^3 k} {(2 \pi)^3} \langle G_a(k; i \varepsilon ) G_b(-k;i \omega - i \varepsilon ) \rangle \label{n2} \end{equation} whereby the $k$-integration is understood to be carried out after the configurational average has been performed, since before that both resolvents would depend non-trivially on two momentum variables. We consider our system at zero temperature and follow partly the method used in Ref. \onlinecite{bi15} for our calculations. At $T=0$, the polarization can be obtained as the energy convolution around the conduction band branch cut of the $k$-summed vertex corrected two particle function $\displaystyle{K(z_1,z_2) = \int \frac {d^3 k} {(2 \pi)^3} \langle G_a(k;z_1 ) G_b(-k;z_2 ) \rangle}$, such that \begin{equation} \Pi_r( \omega ) = \oint_C K(z, \omega + i \delta - z) dz \label{n3} \end{equation} where we have taken over the following definitions from Ref. \onlinecite{bi15}: \begin{equation} K(z_1,z_2) = \frac {R(z_1,z_2)} {1-\Lambda(z_1,z_2)R(z_1,z_2)} \label{n4} \end{equation} where $ \Lambda(z_1,z_2) $ is the usual CPA vertex from (\ref{ca17}) and \begin{equation} R(z_1,z_2) = \int \frac {d^3 k} {(2 \pi )^3} \bar G_a(k,z_1) \bar G_b(-k,z_2) \label{n5} \end{equation} is the average-decoupled two particle function. Assuming that the conduction and valence band dispersion relations exhibit a similar shape such that they scale proportionally \begin{equation} \frac { \varepsilon_a(k)} {w_a} = \mp \frac { \varepsilon_b(k)} {w_b} \label{n6} \end{equation} (\ref{n5}) can be shown to simplify to \begin{equation} R(z_1,z_2) = \frac {w_a F_a(z_1) \pm w_b F_b(z_2)} {w_a[z_2 - \Sigma_b(z_2)] \pm w_b[z_1 - \Sigma_a(z_1)]} \label{n7} \end{equation} where $F_\lambda (z)$ are the site diagonal single particle functions first introduced in connection with (\ref{b25}). As usual we assume that the effective mass of an electron in the conduction band is positive and that of a hole in the valence band is negative. Accordingly we have chosen the upper choice of signs in (\ref{n6}) and (\ref{n7}) for our calculations. To be able to analyze the obtained results with regard to the effect of the the inclusion of vertex corrections we first consider qualitatively the features that would be expected from the transition process represented by the energy convolution in (\ref{n3}) in an intermediate regime of disorder, if the configurational average in the two particle function is decoupled and effectively only single particle properties are employed. This would correspond to replacing $K$ from (\ref{n4}) by $R$ from (\ref{n5})in (\ref{n3}). We assume for now that the concentration be about 0.5 and the bands have just split by a notable amount. With the semi-elliptic bands used, the transition process can be represented as shown in Fig. \ref{FIG2}. The disorder strengths give approximately the distance between the centers of the single bands. The convolution of two separated finite bands, occurring in $R$, would yield a set of finite bands in the joint density of states (DOS) whose width is the sum of the widths of the contributing components. Two cases are considered where the band offsets of the $A-$ and $B$-components of the alloy are in equal or opposite directions corresponding to parallel or anti-parallel disorder. In the case of parallel disorder, this would amount to the $A_b \rightarrow A_a$ and $B_b \rightarrow B_a$ -transitions lying in the center of the joint DOS, framed by the contributions from the $B_b \rightarrow A_a$ and $A_b \rightarrow B_a$ -transitions upper and lower end respectively as shown in Fig. \ref{FIG3}. At $c=0.5$ these regions would have relative distribution of weighted states of $1:2:1$ from lower end : center : higher end. In the case of anti-parallel disorder the picture should be similar with the only difference that the spectrum is turned inside-out with the $B_b \rightarrow B_a$ and $A_b \rightarrow A_b$ components on the top and the bottom end of the joint DOS and the mixed transitions in the center, again with a distribution of $1:2:1$. The calculations in the previous section for the split band case strongly suggest that the vertex corrections will increasingly suppress the cross transitions as the disorder strength is increased, which is verified in our numerical results. Indeed, our results show that this suppression is already displayed quite strongly in an intermediate disorder range, i.e. in a regime where the single bands just begin to split. The appended plots in Figs. \ref{FIG4} and \ref{FIG5} show cumulative absorption spectra calculated from (\ref{n3}) as well as their single and double weighted components for parallel and anti-parallel disorder of various strengths, covering both the joint and the split band regime. The conduction band half-width is normalized to unity and the valence band half-width is taken to be $0.8$. The concentration of impurities is fixed to $0.35$ in order to study the high-concentration behavior rather than dilute impurity effects. In the transitional region when the disorder strengths start to exceed the single particle half-bandwidths $ \mid V_{ \lambda } / w_{ \lambda } \mid \geq 1$ and the conduction and valence bands start to split we observe the following behavior: In the case of parallel disorder shown in Fig. \ref{FIG4} the spectrum starts to exhibit a discontinuity in its derivatives at the flanks accounting for a pair of mixed components splitting off sideways from the main contribution. At the transition point the contributions of the flanks relative to the central bulk part are approximately distributed in a ratio of $1:14:1$. In the case of anti-parallel disorder shown in Fig. \ref{FIG5} the suppression of the mixed transitions is even stronger so that as soon as the single bands split the crossed transitions cancel out entirely within the numerical accuracy and the joint density of states starts to exhibit a gap between two separate contributions which mainly consist of $A-$ and $B$-transitions, respectively. In a regime of strong disorder, further beyond the splitting of the single bands as shown in the last plots of Figs. \ref{FIG4} and \ref{FIG5}, one finds that over large regions the single and uniformly double weighted contributions coincide almost exactly, implying that the spectrum is built almost entirely from the total diagonal element of the two particle function, which we had calculated at the end of section IV. The total diagonal element can be obtained as an independent $k$-sum over the two single particle resolvents involved. This situation represents a breakdown of the $k$-selection rule which holds in pure media. If the splitting of $\Pi$ into single weighted components $\Pi = \Pi^A + \Pi^B $ is considered and the result is compared in appearance with the splitting into components of the site diagonal single particle function \cite{bi3} \begin{equation} F_r( \omega ) = \int \frac {d^3 k} {(2 \pi)^3} \bar G(k, \omega + i \delta) \end{equation} one finds that the single and the two particle behavior appear to be strikingly similar in the case of anti-parallel disorder, as can be seen if the plots for the imaginary part of the single particle function shown in Figs. \ref{FIG6} (a) and (b) are compared to the ones for the corresponding parameters in the two particle case of Fig. \ref{FIG5} (third and fourth plot from the front). In a system without disorder this similarity is evident if there are no further local interactions in the problem, since the non-interacting two particle motion decouples into a center of mass and a relative co-ordinate and while the center of mass motion can be set to zero, the relative one can be mapped onto a single particle co-ordinate. Upon the addition of the disorder, this decoupling fails to work and it can only be regained by using an appropriate configurational averaging procedure. However, if in the presence of disorder an average is only performed on a single particle level, thus omitting average induced two particle correlations the reduction obviously fails to work, as is seen through comparing the plots of Fig. \ref{FIG7} showing a spectrum calculated without the vertex corrections with the ones for the corresponding parameter values of Fig. \ref{FIG5}, which properly include these corrections. The results show, that in regimes of intermediate and strong disorder the influence of the vertex corrections is very substantial. In comparison to the single weighted (two fold) splitting the double weighted (three fold) splitting exhibits a rather curious behavior. Even though the components $-$Im$\Pi^{AA}$ and $-$Im$\Pi^{BB}$ lie underneath their single weighted complements $-$Im$\Pi^{A}$ and $-$Im$\Pi^{B}$ in some parts of the spectra, which one might expect to happen globally on first thought, they either coincide with them or even exceed them in other parts -- sometimes to such an extent that they reach beyond the cumulative function. However, it has to be noted, that these components, like the unweighted function, are always uniformly positive in sign, and therefore exhibit the correct analytic behavior that a function defined on this footing has to satisfy. It is required that these components be positive definite, because the net absorption in the medium must always be positive unless the system is excited out of equilibrium, which we do not consider here. In the preceding discussion we have assumed for convenience that the optical matrix elements between states of different components of the alloy are equal. If different optical cross sections (matrix elements) $\mu^{A/B}$ are distinguished for $A-$ and $B$-atoms the weighted contributions to yield the integrand of (\ref{n3}) would sum as \begin{equation} \tilde K= (\mu^A)^2 K^{AA} + 2\mu^A \mu^B K^{AB} + (\mu^B)^2 K^{BB} \label{n8}. \end{equation} This shows that it would be possible to observe one of the functions $\Pi^{AA}$ or $\Pi^{BB}$ predominantly if either $\mu^A$ or $\mu^B$ happens to be much larger than the other. The mixed function $\Pi^{AB}$, however, will never be a separately observable quantity in a general case, no matter how the cross sections $\mu^A$ and $\mu^B$ scale relatively and therefore not so rigid criteria for its analyticity apply as for the uniformly double weighted functions. \section{Discussion and Conclusion} In the previous sections we have obtained expressions for a wide class of weighted two particle Green's functions. The large choice for possible weightings is substantially reduced as restrictions are made to functions which would be useful in linear response theory. In both cases which are discussed for this kind of application, only five different weighted functions remain of which only two are genuinely independent. The structure of the weighting process is closely related to that derived for the single particle theory with the main difference that now the weights also depend significantly on the CPA vertex corrections. The calculation for the split band limit, the domain in which the CPA is superior to most of the other theories of disordered systems, gives a direct insight into how the properly weighted CPA extracts the correct limiting behavior from different possible physical processes. Some care is needed in interpreting the precise physical meaning of the weights, since they are obtained for the averaged functions, which describe the disordered medium as effectively homogeneous. The concept of the propagation of a particle between sites of different components is therefore lost in the effective medium as a consequence of averaging and the initial exclusion of specific propagation paths in the unaveraged function leads to the effective weights. These weights simply account for the average partition in probability for the simultaneous propagation of the particles between partly or completely specified site types at a given pair of energies $z_1$ and $z_2$. Our numerical results show the general importance of the inclusion of vertex corrections into a properly self consistent two particle formalism. We managed to visualize that, as a consequence of the inclusion of these average induced two particle correlations, the center of mass and relative motion of the two particle system effectively decouple to a large extent. We believe that the general method developed here will find applications in various situations where two particle motion is studied in a disordered medium. The effect of alloying on the electronic susceptibility and hence for example on the Rudermann-Kittel interaction has already been mentioned. In particular we believe it can be extended for use in systems where the two particles have a direct interaction, such as the Coulomb interaction between carriers occurring in excitons in alloyed systems. In such a case the static correlations between particles, created by the disorder and accounted for by the vertex corrections, and the dynamic correlations introduced through the carrier-carrier interaction, create additional static-dynamic correlations. Moreover it may be possible that the underlying disorder of the system gives rise to an induced disorder to the carrier-carrier interaction itself. Both of these effects can be treated within the method developed here and will be addressed in a forthcoming publication. \section{Acknowledgements} The authors would like to thank Mr. Carsten Heide at Oxford for reading the manuscript and making useful suggestions.
\section{Introduction} To determine the proton charge radius with a percent accuracy from the value of the hydrogen Lamb shift, the latter should be known both experimentally and theoretically with a precision of one kHz. Recently completed calculation of the order $m\mbox{$\alpha$}^2(\mbox{$Z\alpha$})^5$ corrections \cite{E} leaves among the effects of possible phenomenological interest the pure recoil correction of the order $m^2(\mbox{$Z\alpha$})^6/M$ arising due to interference between the nucleus' recoil and relativistic effects in the motion of the electron. The present paper is devoted to the calculation of this correction for an arbitrary state of the hydrogen atom. Recently this correction for $P$ states was found \cite{p}. In those states, as well as in all states with nonzero angular momenta, the correction proves to be saturated by a contribution coming from the atomic scale. Hence one can use there the standard quantum mechanical perturbation theory for the effective operators describing relativistic effects. Matrix elements of the effective operators arising in the perturbation theory converge at small distances thus testifying {\it a posteriori} that the used 'nonrelativistic' approach is correct at the given order of \mbox{$\alpha$}\ for states with nonzero $l$. An attempt to apply the same approach to $S$ states, whose wavefunctions do not vanish at the origin, leads to matrix elements diverging at small distances. In fact, among the effective operators one finds those depending on $r$ as $r^{-3}$ and even $r^{-4}$ \cite{p}. As for the latter, for $S$ states the operator $r^{-4}$ is equivalent (modulo a nonsingular operator) to the sum of operators with the radial dependence $r^{-3}$ and $\delta(\mbox{$\vec{r}$})/r$. It was shown in Ref.\cite{FKMY} that logarithmically divergent contributions are mutually cancelled. This cancellation means that for the states with vanishing angular momentum, the correction we discuss splits naturally into two contributions -- those of large and small distances -- each gaining its value in its own scale. To calculate the former, one can use again the nonrelativistic approach, whereas the short distance contribution, residing in the Compton wavelength order scale, calls for a true relativistic approach. The closed expression for the first recoil correction to an energy of the relativistic electron moving in the Coulomb field is outlined in Sec.2. This expression is used in Sec.3 for the evaluation of the long-distance contribution. It proves that the relativistic approach is more efficient even at the atomic scale. The contribution of short distances is found in Sec.4 employing the Feynman gauge. Sec.5 is devoted to checks of the obtained results. The long-distance contribution is recalculated there using the nonrelativistic approach, while the short-distance one is found in the Coulomb gauge. Finally, in Sec.6, we give the numerical values for the energy shifts and compare results of the present work with those obtained earlier in \cite{p,PG}. Throughout the paper the relativistic units $\hbar=c=1$ are used. Since we do not discuss radiative corrections, $Z$ is also set equal to unity. \section{Methods of Calculation} One of the perturbation schemes we use at the present work starts from the Schr\"odinger equation in the Coulomb field, when both particles are considered as nonrelativistic in the zeroth approximation. To account for relativistic effects, a kind of the operator product expansion is built by calculation of scattering amplitudes for free relativistic particles. Thus arising effective operators are then the subject for the ordinary perturbation theory of the nonrelativistic quantum mechanics. This approach is rather well suited for long-distance contributions, which are due to effective operators saturated by nonrelativistic region and having non-local kernels. Formerly it was used in calculations of i) logarithmic in \mbox{$\alpha$}\ corrections to the spectrum of the two-body system \cite{PhScr}, ii) the order $m\mbox{$\alpha$}^6$ corrections to the positronium $P$ levels \cite{pos}, and iii) the order $m^2\mbox{$\alpha$}^6/M$ corrections to the hydrogen $P$ levels \cite{p}. Unfortunately, this approach becomes very tedious being applied to short-distance contributions, when effective operators with local kernels are represented by a number of diagrams. An alternative approach deals with relativistic light particle (electron) moving in the field generated by the slow heavy one (nucleus). In the zeroth approximation, the heavy particle holds still being a source of the Coulomb field. Wavefunction of the system reduces to that of the light particle satisfying the Dirac equation. To first order in the heavy particle's inverse mass, the perturbation operator coincides with its nonrelativistic Hamiltonian: \begin{equation}\label{Hh} V = \fr{\left ( \vec{P} - |e| \vec{A}(\vec{R}) \right ) ^2}{2M}. \end{equation} Here $\vec{P}$ is the operator of a nucleus' momentum. The vector potential $\vec{A}$ acts at the nucleus' site. Unfortunately, one cannot calculate an energy shift induced by the perturbation (\ref{Hh}) straightforwardly, i.e. taking merely its average. In fact, the operator (\ref{Hh}) depends on the nucleus' dynamical variables while the argument of the unperturbed wavefunction is a position or momentum of the electron. To overcome this difficulty we use the gauge invariance of observables in the quantum electrodynamics \cite{Y}. Being reexpressed in terms of electron's variables, the average value of (\ref{Hh}) should be retained gauge invariant. The new form of the average is now nearly evident: \begin{equation}\label{main} \mbox{$\Delta E$}_{rec} = - \fr{1}{M} \int \fr{d\mbox{$\omega$}}{2\pi i} \left\langle \left ( \mbox{$\vec{p}$} - \mbox{$\vec{D}$} \right ) G_{E+\mbox{$\omega$}} \left ( \mbox{$\vec{p}$} - \mbox{$\vec{D}$} \right ) \right\rangle. \end{equation} Here \mbox{$\vec{p}$}\ is the electron momentum operator, \mbox{$\vec{D}$}\ is the integral operator describing the transverse quantum exchange. It has the kernel\footnote{In what follows we will write often a kernel rather than an appropriate operator for the sake of brevity.} \[ \fr{4\pi \mbox{$\alpha$} \mbox{$\vec{\alpha}$}_k }{k^2 - \mbox{$\omega$}^2}, \;\;\;\;\; \mbox{$\vec{\alpha}$}_k \equiv \mbox{$\vec{\alpha}$} - \fr{\mbox{$\vec{k}$} (\mbox{$\vec{\alpha}$}\mbox{$\vec{k}$})}{k^2}. \] In (\ref{main}), $G$ is the Green's function for the Dirac equation in the Coulomb field, the average is taken over a Dirac-Coulomb eigenstate with an energy $E$. Actually, the "seagull" part of (\ref{main}), that of the second order in \mbox{$\vec{D}$}, emerges naturally as a counterpart of the $\vec{A}^2$ term from (\ref{Hh}) when we take the expectation value of this term over fluctuations of the electromagnetic field. All the other terms provide the invariance of (\ref{main}) with respect to the gauge transformation, \begin{equation} \psi \mbox{$\rightarrow$} \exp(i\phi(\mbox{$\vec{r}$}))\psi,\;\;\;\;\;\;\mbox{$\vec{D}$} \mbox{$\rightarrow$} \mbox{$\vec{D}$} + i[\mbox{$\vec{p}$},\phi]. \end{equation} One can easily convince oneself that just the same result can be obtained from the formula (11) of \cite{Y} with the help of the Dirac equation. The first attempt to obtain the relativistic expression for the recoil correction to hydrogen energy levels was undertaken in Ref.\cite{Br}. Complete expressions for various contributions in the Coulomb gauge were originally derived in the framework of the quasipotential approach in Ref.\cite{Sh}. The sum of those contributions can be convinced to reduce to the right-hand side of (\ref{main}). \section{Long-Distance Contribution} Present section is devoted to the calculation of the contribution to the energy shift which is saturated by the atomic scale and can thus be called the long-distance one. To check the results, two approaches described above have been applied in parallel. Here we describe in detail how the second, relativistic, approach works. The procedure of comparison with the results of the more cumbersome nonrelativistic approach will be postponed until the Section 5. \subsection{Pure Coulomb Contribution} In the relativistic approach the pure Coulomb contribution, \begin{equation} \mbox{$\Delta E$}_C = - \fr{1}{M} \int \fr{d\mbox{$\omega$}}{2\pi i} \left\langle \mbox{$\vec{p}$} G_{E+\mbox{$\omega$}} \mbox{$\vec{p}$} \right\rangle = \fr{1}{2M} \left\langle \mbox{$\vec{p}$} \left ( \Lambda_+ - \Lambda_- \right ) \mbox{$\vec{p}$} \right\rangle, \end{equation} can naturally be represented as the sum of two terms \cite{Br}, \begin{equation}\label{C} \mbox{$\Delta E$}_C = \left\langle \fr{p^2}{2M} \right\rangle - \fr{1}{M} \left\langle \mbox{$\vec{p}$} \Lambda_- \mbox{$\vec{p}$} \right\rangle. \end{equation} Here $\Lambda_+$ and $\Lambda_-$ are the projection operators to sets of positive- and negative-energy Dirac-Coulomb eigenstates respectively. With the aid of the Dirac equation the mean value of $p^2/2M$ can readily be reexpressed in the following form \cite{Sh}: \begin{equation}\label{kin} \left\langle \fr{p^2}{2M} \right\rangle = \fr{m^2-E^2}{2M} + \fr{m^2}{2M} \left\langle 2 \left ( \fr{E}{m} -\mbox{$\beta$} \right ) \fr{\mbox{$\alpha$}}{r} + \fr{\mbox{$\alpha$}^2}{r^2} \right\rangle. \end{equation} As for the second term in (\ref{C}), responsible for virtual transitions into negative-energy states, the simple analysis shows that it doesn't contribute to the order of interest at the atomic scale. Actually, the trivial power counting on the right-hand side of the obvious inequality, \begin{equation}\label{in} |\left\langle \mbox{$\vec{p}$} \Lambda_- \mbox{$\vec{p}$} \right\rangle| < \left|\fr{1}{4m^2}\left\langle [\mbox{$\vec{p}$},C] \Lambda_- [\mbox{$\vec{p}$},C] \right\rangle \right|, \end{equation} where $C$ is the Coulomb potential, shows that at the atomic scale, the product of commutators is already of the sixth order in \mbox{$\alpha$}, so that the projector and the wavefunctions can sufficiently be replaced by their nonrelativistic counterparts. Since there is no negative-energy states in the nonrelativistic approximation, the atomic scale contribution to the initial average also vanishes in the order we consider. \subsection{Magnetic Contribution} After performing the integration over \mbox{$\omega$}, the expression for the single transverse, or magnetic, contribution, \begin{equation}\label{M} \mbox{$\Delta E$}_M = \fr{1}{M} \int \fr{d\mbox{$\omega$}}{2\pi i} \left\langle \mbox{$\vec{p}$} G \mbox{$\vec{D}$} + \mbox{$\vec{D}$} G \mbox{$\vec{p}$} \right\rangle, \end{equation} turns into \begin{equation}\label{Mld} \mbox{$\Delta E$}_M = -\fr{\mbox{$\alpha$}}{M}\mbox{Re} \left\langle \mbox{$\vec{p}$} \left ( \sum_+ \fr{|m\rangle\langle m|}{k+E_m-E} - \sum_-\fr{|m\rangle\langle m|}{E-E_m+k} \right ) \fr{4\pi \mbox{$\vec{\alpha}$}_k}{k} \right\rangle, \end{equation} where $\sum_+$ denotes the sum over discrete levels plus the integral over positive-energy part of the continuous spectrum, while $\sum_-$ stands for the integral over negative-energy continuum. For the transverse photon momenta in the atomic region, $k \sim m\mbox{$\alpha$}$, one can expand the first term in (\ref{Mld}) to the power series in the ratio $(E-E_m)/k$. To zeroth order (in the approximation of the instant exchange), we have: \begin{equation}\label{M0} -\fr{\mbox{$\alpha$}}{M}\mbox{Re} \left\langle \fr{4\pi \mbox{$\vec{\alpha}$}_k}{k^2}\Lambda_+ \mbox{$\vec{p}$} \right\rangle = - \fr{m^2}{2M} \left\langle 2 \left ( \fr{E}{m} -\mbox{$\beta$} \right ) \fr{\mbox{$\alpha$}}{r} + \fr{\mbox{$\alpha$}^2}{r^2} \right\rangle +\fr{\mbox{$\alpha$}}{M}\mbox{Re} \left\langle \fr{4\pi \mbox{$\vec{\alpha}$}_k}{k^2}\Lambda_- \mbox{$\vec{p}$} \right\rangle. \end{equation} The sum of the first term and (\ref{kin}) has very simple form \cite{Sh}, \begin{equation}\label{Inst} \fr{m^2-E^2}{2M} = \fr{m^2\mbox{$\alpha$}^2}{2MN^2}, \end{equation} where the standard notations for the Dirac-Coulomb problem are used, \[ N = \sqrt{(\gamma+n_r)^2 + \mbox{$\alpha$}^2}, \;\;\;\;\gamma = \sqrt{\mbox{$\kappa$}^2-\mbox{$\alpha$}^2}, \] $n_r$ is the radial quantum number, $\mbox{$\kappa$} = - 1 - \vec{\sigma}\vec{l}$. Notice that (\ref{Inst}) reduces to the lowest order result for the states with $n_r=0$ only. As far as we are seeking only for corrections of the even order in \mbox{$\alpha$}, the next term of the expansion to be considered is \begin{equation}\label{M2} \mbox{$\Delta E$}_{ret}=-\fr{\mbox{$\alpha$}}{M}\mbox{Re} \left\langle \mbox{$\vec{p}$} \sum_+ (E_m-E)^2 |m\rangle\langle m| \fr{4\pi \mbox{$\vec{\alpha}$}_k}{k^4} \right\rangle = -\fr{\mbox{$\alpha$}}{M}\mbox{Re} \left\langle \left[ H, \left[ H, \mbox{$\vec{p}$} \right] \right] \Lambda_+ \fr{4\pi \mbox{$\vec{\alpha}$}_k}{k^4} \right\rangle, \end{equation} where $H = \mbox{$\vec{\alpha}$}\mbox{$\vec{p}$} + \mbox{$\beta$} m + C$ is the Dirac Hamiltonian in the Coulomb field. This term describes the effect of retardation. Implying the corresponding operator by its kernel, we have \[ [H,\mbox{$\vec{p}$}] = \mbox{$\alpha$} \fr{4\pi\mbox{$\vec{k}$}}{k^2}, \;\;\;\;\; \left[H,[H,\mbox{$\vec{p}$}]\right] = \mbox{$\alpha$} \fr{4\pi\mbox{$\vec{k}$}(\mbox{$\vec{\alpha}$}\mbox{$\vec{k}$})}{k^2}. \] To the lowest nontrivial order, matrix elements of \mbox{$\vec{\alpha}$}'s over positive-energy states can be replaced by the appropriate Pauli currents: \begin{equation}\label{pret} \mbox{$\Delta E$}_{ret}\approx - \fr{\mbox{$\alpha$}^2}{M} \left\langle \fr{4\pi\mbox{$\vec{k'}$}}{k'^2} \fr{2\mbox{$\vec{p}\,'$}\mbox{$\vec{k'}$}-k'^2}{2m}\;\fr{4\pi}{k^4} \fr{2\mbox{$\vec{p}$}_k+i\vec{\sigma}\times\mbox{$\vec{k}$}}{2m} \right\rangle. \end{equation} Here $\mbox{$\vec{k'}$} = \vec{q}-\mbox{$\vec{k}$}$, $\vec{q}=\mbox{$\vec{p}\,'$}-\mbox{$\vec{p}$}$, while \mbox{$\vec{p}$}\ and \mbox{$\vec{p}\,'$}\ are the arguments of the wavefunction and its conjugated respectively. Being converted to the spatial representation, the average above equals \begin{equation}\label{ret} \mbox{$\Delta E$}_{ret} = \fr{\mbox{$\alpha$}^2}{4m^2M} \left\langle -2\mbox{$\vec{p}$}\fr{1}{r^2}\mbox{$\vec{p}$} + \fr{7\vec{l}^2 + 2\vec{\sigma}\vec{l}}{2r^4} \right\rangle. \end{equation} Strictly speaking, in (\ref{pret}), the integral over \mbox{$\vec{k}$}\ has an infrared divergent part. It is omitted from (\ref{ret}) since the photon momenta $k\sim m\mbox{$\alpha$}^2$ contribute to the previous order correction. To make sure that this is correct, one can regularize the divergency supplying the photon with a mass $\lambda$ such that $m\mbox{$\alpha$}^2 \ll \lambda \ll m\mbox{$\alpha$}$. The term proportional to $1/\lambda$ and omitted from (\ref{ret}) can be easily checked to cancel the respective term in the difference between (\ref{Mld}) and the expression obtained from (\ref{Mld}) by the replacement $k \mbox{$\rightarrow$} \sqrt{k^2+\lambda^2}$. On the other hand, just this difference determines the low-energy contribution to the order $m\mbox{$\alpha$}^5/M$ correction. The last of contributions due to the single transverse exchange is generated by virtual transitions into negative-energy states and is covered by the last terms in (\ref{Mld}) and (\ref{M0}). The inequality similar to (\ref{in}) shows that the nonrelativistic expansion of the last term in (\ref{Mld}) starts with the seventh power of \mbox{$\alpha$}. As for the negative-energy contribution to (\ref{M0}), to the lowest nontrivial order it reduces to \begin{equation}\label{M-mom} - \fr{\mbox{$\alpha$}^2}{2mM} \left\langle \fr{4\pi \mbox{$\vec{\alpha}$}_{k'}}{k'^2} \lambda_-(\mbox{$\vec{p}$}+\mbox{$\vec{k}$}) \fr{4\pi \mbox{$\vec{k}$}}{k^2} \right\rangle \approx \fr{\mbox{$\alpha$}^2}{4m^2M} \left\langle \fr{4\pi \mbox{$\vec{\alpha}$}_{k'}\mbox{$\vec{k}$}}{k'^2} \fr{4\pi \mbox{$\vec{\alpha}$} \mbox{$\vec{k}$}}{k^2} \right\rangle, \end{equation} yielding in the spatial representation \begin{equation}\label{M-} \mbox{$\Delta E$}_{M-} = - \fr{\mbox{$\alpha$}^2}{4m^2M} \left\langle \fr{1}{r^4} - \fr{4\pi\delta(\mbox{$\vec{r}$})}{r} \right\rangle. \end{equation} Taken over $S$ states, this average is logarithmically divergent at small distances (linear divergencies cancel each other). An ultraviolet divergency will be discussed later for the {\em total} long-distance contribution to the $S$ level shift. In fact, due to the gauge dependence of an individual contribution (e. g., (\ref{M-})), its divergent part alone has no physical meaning. \subsection{Seagull Contribution} Again, taking the integral over \mbox{$\omega$}\ in the expression for the double transverse, or seagull, contribution, \begin{equation}\label{S} \mbox{$\Delta E$}_S = - \fr{1}{M} \int \fr{d\mbox{$\omega$}}{2\pi i} \left\langle \mbox{$\vec{D}$} G_{E+\mbox{$\omega$}} \mbox{$\vec{D}$} \right\rangle, \end{equation} we obtain \begin{equation}\label{Sld} \mbox{$\Delta E$}_S = \fr{\mbox{$\alpha$}^2}{2M} \left\langle \fr{4\pi \mbox{$\vec{\alpha}$}_{k'}}{k'} \sum_+ \fr{|m\rangle\langle m|}{(E_m-E+k')(E_m-E+k)} \left ( 1 + \fr{E_m-E}{k'+k}\right ) \fr{4\pi \mbox{$\vec{\alpha}$}_k}{k} + \cdots \right\rangle, \end{equation} where the ellipsis stands for the negative-energy part which differs from the positive-energy one by the overall sign and signs before $k$ and $k'$. One can easily check that the linear in $k/2m$ terms in the expansion of the negative-energy part cancel each other. But just these terms at the atomic scale could produce the energy correction of the necessary order. Hence it remains to consider the positive-energy part explicitly written in (\ref{Sld}). In the leading nonrelativistic approximation, \begin{equation} \mbox{$\Delta E$}_{S+} = \fr{\mbox{$\alpha$}^2}{2M}\left\langle\fr{4\pi\mbox{$\vec{\alpha}$}_{k'}}{k'^2}\Lambda_+ \fr{4\pi\mbox{$\vec{\alpha}$}_k}{k^2}\right\rangle, \end{equation} we again replace matrix elements of \mbox{$\vec{\alpha}$}'s over positive-energy states by the Pauli currents, \begin{equation} \mbox{$\Delta E$}_{S+} = \fr{\mbox{$\alpha$}^2}{2M}\left\langle\fr{4\pi}{k'^2}\fr{2\mbox{$\vec{p}\,'$}_{k'}+i\vec{\sigma} \times\mbox{$\vec{k'}$}}{2m}\fr{4\pi}{k^2}\fr{2\mbox{$\vec{p}$}_k+i\vec{\sigma} \times\mbox{$\vec{k}$}}{2m} \right\rangle, \end{equation} and perform the Fourier transformation to obtain \begin{equation}\label{S+} \mbox{$\Delta E$}_{S+} = \fr{\mbox{$\alpha$}^2}{4m^2M} \left\langle 2\mbox{$\vec{p}$}\fr{1}{r^2}\mbox{$\vec{p}$} + \fr{1}{r^4} - \fr{3\vec{l}^2 + 2\vec{\sigma}\vec{l}}{2r^4} \right\rangle. \end{equation} \subsection{Total Long-Distance Contribution} Summing up (\ref{ret}), (\ref{M-}) and (\ref{S+}) we arrive at \begin{equation}\label{sum} \fr{\mbox{$\alpha$}^2}{4m^2M} \left\langle 2 \fr{\vec{l}^2}{r^4} + \fr{4\pi\delta(\mbox{$\vec{r}$})}{r} \right\rangle. \end{equation} Expanding (\ref{Inst}) to the power series in $\mbox{$\alpha$}^2$ and evaluating the average in (\ref{sum}) we obtain the long-distance contribution for a state with nonzero $l$: \begin{eqnarray}\label{ld} \mbox{$\Delta E$}_{l>0} &=& \fr{m^2\mbox{$\alpha$}^6}{Mn^3} \left\{ \fr{1}{8|\mbox{$\kappa$}|^3} + \fr{6\mbox{$\kappa$}}{|\mbox{$\kappa$}|(4\mbox{$\kappa$}^2-1)(2\mbox{$\kappa$}+3)} + \fr{3}{8n\mbox{$\kappa$}^2} \right. \nonumber \\ &&\left. - \fr{1}{n^2|\mbox{$\kappa$}|} \left ( 1 + 2\fr{\mbox{$\kappa$}^2 (\mbox{$\kappa$}+1)}{(4\mbox{$\kappa$}^2-1)(2\mbox{$\kappa$}+3)} \right ) + \fr{1}{2n^3} \right\}. \end{eqnarray} For $l=1$ it reproduces the result of the paper \cite{p}. As previously mentioned, the effective operators of the order under consideration contain singularities insufficient to compensate the vanishing of a wavefunction with nonzero $l$ at $r\mbox{$\rightarrow$} 0$. That is why (\ref{ld}) is the total sought-for correction to $l>0$ levels. For $S$ states the first term in (\ref{sum}) evidently vanishes due to the angular momentum operator $\vec{l}$ annihilating their wavefunctions. It is interesting to note that a na\"{\i}ve generalization of the result for states with nonzero $l$ to $S$ ones leads to the error -- the vanishing of the angular average is compensated by the linear divergency of the radial one. As for the second term, which formally contains the linear divergency, it is just a remnant of the short-distance contribution to the previous (fifth in \mbox{$\alpha$}) order correction. To make sure that nothing is lost in the sixth order, let us regularize the ultraviolet divergency by subtraction of the potential generated by the massive transverse exchange (the photon mass $\lambda \gg m\mbox{$\alpha$}$), from the potential of the ordinary transverse exchange entering (\ref{M-mom}): \begin{equation}\label{UVreg} \fr{4\pi \mbox{$\vec{\alpha}$}_{k'}}{k'^2} \;\; \mbox{$\rightarrow$} \;\; \fr{4\pi \mbox{$\vec{\alpha}$}_{k'}}{k'^2} - \fr{4\pi \mbox{$\vec{\alpha}$}_{k'}}{k'^2 + \lambda^2}. \end{equation} By going to the spatial representation we obtain the regularized version of the singular operator: \[ \fr{4\pi\delta(\mbox{$\vec{r}$})}{r} \;\; \mbox{$\rightarrow$} \;\; \fr{2\lambda}{3} 4\pi\delta(\mbox{$\vec{r}$}). \] Being averaged, it gives the energy correction of the order $m\lambda\mbox{$\alpha$}^5/M$. The latter should be cancelled by the linear in $\lambda$ term in the expansion of the short-distance contribution to the order $m^2\mbox{$\alpha$}^5/M$ correction, calculated with the massive propagator of the transverse quantum (actually the expansion parameter is $\lambda/m \ll 1$). Along with linear in $\lambda/m$ correction, one could expect the correction linear in $\mbox{$\alpha$}=m\mbox{$\alpha$}/m$. However the expansion parameter at large distances is $(p/m)^2 \sim \mbox{$\alpha$}^2$ so that the operator we discuss does not contribute to the order of interest. On the other hand, a linear in \mbox{$\alpha$}\ correction to a local ($\propto \delta(\mbox{$\vec{r}$})$) operator can arise as an ordinary radiative one. In this case the correction is completely saturated by small distances. The next section is devoted to the calculation of such corrections. So, in $S$ states the long-distance contribution is exhausted by the $m^2\mbox{$\alpha$}^6/M$ term from the expansion of (\ref{Inst}), \begin{equation}\label{ldS} \mbox{$\Delta E$}_{l=0}^{ld} = \fr{m^2\mbox{$\alpha$}^6}{2Mn^3} \left ( \fr{1}{4} + \fr{3}{4n} - \fr{2}{n^2} + \fr{1}{n^3} \right ). \end{equation} It vanishes in the ground state only. \section{Short-Distance Contribution} As we have mentioned in the Introduction, the long-distance contribution for $S$ states is supplied by a short-distance one residing at scales of the Compton wavelength order. Since two contributions are well separated, each of them is gauge invariant, so that evaluating the short-distance one we can use a different, more appropriate gauge. Rather naturally the mostly convenient gauge is the Feynman one. The main formula (\ref{main}) can be rewritten in this gauge by application of the Dirac equation or directly from the eq.(11) of \cite{Y}: \begin{equation}\label{fey} \mbox{$\Delta E$}_{tot} = - \fr{\mbox{$\alpha$}^2}{M} \int \fr{d\mbox{$\omega$}}{2\pi i} \left\langle \fr{4\pi}{k'^2 + \lambda^2 - \mbox{$\omega$}^2} \left ( \mbox{$\vec{\alpha}$} + \fr{\mbox{$\vec{k'}$}}{\mbox{$\omega$}} \right ) G_{E+\mbox{$\omega$}} \left ( \mbox{$\vec{\alpha}$} - \fr{\mbox{$\vec{k}$}}{\mbox{$\omega$}} \right ) \fr{4\pi}{k^2 + \lambda^2 - \mbox{$\omega$}^2} \right\rangle. \end{equation} Momenta of the photons are assumed to flow both from the nucleus to the electron. The photon mass $\lambda$ is introduced to establish control over infrared divergences reminiscent of lower-order and long-distance contributions. Those divergencies arise in the process of the approximate evaluation of the integrals in (\ref{fey}). Taking the wavefunctions at the origin and replacing the Green's function by the first term of its expansion in the Coulomb field, we have \begin{eqnarray} \mbox{$\Delta E$}_G = \fr{\mbox{$\alpha$}^3\psi^2}{M}\int \fr{d\mbox{$\omega$}}{2\pi i}\left\langle\fr{4\pi}{p'^2- \sqrt{\;}^2}\left (\mbox{$\vec{\alpha}$}-\fr{\mbox{$\vec{p}$}\,'}{\mbox{$\omega$}}\right ) \fr{m+\mbox{$\omega$}+\mbox{$\beta$} m+\mbox{$\vec{\alpha}$}\mbox{$\vec{p}$}\,'}{p'^2-\Omega^2}\fr{4\pi}{q^2}\right.&&\nonumber\\ \left.\fr{m+\mbox{$\omega$}+\mbox{$\beta$} m+\mbox{$\vec{\alpha}$}\mbox{$\vec{p}$}}{p^2-\Omega^2}\left (\mbox{$\vec{\alpha}$}-\fr{\mbox{$\vec{p}$}}{\mbox{$\omega$}}\right ) \fr{4\pi}{p^2-\sqrt{\;}^2}\right\rangle.&& \end{eqnarray} Here $\psi^2\equiv |\psi(0)|^2$, the angle brackets denote integrations over $\mbox{$\vec{p}$}$ and $\mbox{$\vec{p}$}\,'$; $\vec{q} = \mbox{$\vec{p}$}\,'-\mbox{$\vec{p}$}$; and \[ \sqrt{\;}\equiv \sqrt{\mbox{$\omega$}^2-\lambda^2}, \;\;\;\;\; \Omega\equiv\sqrt{2m\mbox{$\omega$}+\mbox{$\omega$}^2}. \] Contrary to the case of large distances, in the deep relativistic region the opposite order of integration is suitable -- first over \mbox{$\vec{p}$}\ and \mbox{$\vec{p}\,'$}, and then over \mbox{$\omega$}. As for the former, it becomes rather trivial after conversion to the spatial representation. Preparatory to such the conversion, it is convenient to express all the scalar products containing different momenta in terms of their squares. Then some of the denominators can be cancelled. At this point we can drop those terms which do not contain $\Omega$ in their denominators. In fact, the only scale leaving in such terms is $\lambda$ so that they cannot produce a short-distance contribution. In the spatial representation the initial two-loop integral with zero external momenta turns into a simple one-dimensional integral over $r$. The contour of the resulting \mbox{$\omega$}-integration encloses the cut between the points $-2m$ and $-\lambda$ in the complex plane. After this last integration we obtain \begin{equation}\label{G} \mbox{$\Delta E$}_G = \fr{\pi\mbox{$\alpha$}^3\psi^2}{Mm}\left ( \fr{1}{\varepsilon} - \fr{8}{3\pi\sqrt{\varepsilon}}\int_1^{\infty}\fr{dx}{\sqrt{x(x^2-1)}} + 4\ln 2 - \fr{5}{2} \right ), \end{equation} where $\varepsilon=\lambda/2m$. If the Green's function is taken to zeroth order in the interaction, we have to use the Dirac equation in order to account for the momentum dependence in the wavefunction: \begin{eqnarray} \mbox{$\Delta E$}_{\psi} = 2\fr{\mbox{$\alpha$}^3\psi^2}{M}\int \fr{d\mbox{$\omega$}}{2\pi i}\left\langle\fr{4\pi}{p'^4} \left ( 2m+\mbox{$\vec{\alpha}$}\mbox{$\vec{p}$}\,'\right )\fr{4\pi}{q^2-\sqrt{\;}^2} \left (\mbox{$\vec{\alpha}$}+\fr{\vec{q}}{\mbox{$\omega$}}\right )\right.&&\nonumber\\ \left.\fr{m+\mbox{$\omega$}+\mbox{$\beta$} m+\mbox{$\vec{\alpha}$}\mbox{$\vec{p}$}}{p^2-\Omega^2}\left (\mbox{$\vec{\alpha}$}-\fr{\mbox{$\vec{p}$}}{\mbox{$\omega$}}\right ) \fr{4\pi}{p^2-\sqrt{\;}^2}\right\rangle.&& \end{eqnarray} Using the same procedure, we take $-r/2$ as the Fourier transform of $4\pi/p^4$. The linearly divergent constant we thus leave aside is actually proportional to $1/\mbox{$\alpha$}$ and contributes to the previous order correction. The result of the integration is \begin{equation}\label{psi} \mbox{$\Delta E$}_{\psi} = \fr{\pi\mbox{$\alpha$}^3\psi^2}{Mm}\left ( -\fr{1}{2\varepsilon^2} + \fr{1}{\varepsilon} - \fr{8}{3\pi\sqrt{\varepsilon}}\int_1^{\infty}\fr{dx}{\sqrt{x(x^2-1)}}\right ). \end{equation} Regulator-dependent terms in (\ref{G}) and (\ref{psi}) arise from the integrals saturated by the region of momenta $p\sim\lambda$ and frequency $\mbox{$\omega$}\sim\lambda$ (or $\sqrt{m\lambda}$) and are thus the remnants of the previous orders corrections or of the long-distance contribution. Truly relativistic contribution comes from the region $p\sim\mbox{$\omega$}\sim m$ and does not depend on the infrared cutoff: \begin{equation}\label{sd} \mbox{$\Delta E$}^{sd} = \fr{m^2\mbox{$\alpha$}^6}{Mn^3}\left ( 4\ln 2 - \fr{5}{2} \right ) \delta_{l0}. \end{equation} It is pertinent to note here that this result is truly short-distance, i. e. it does not contain hidden long-distance contributions, which na\"{\i}vely could arise due to cancellation of the same nonzero powers of $\lambda$ from numerator and denominator -- all positive powers of the photon mass were dropped out in the process of calculation. On the other hand, an emergence of such contributions would be self-contradictory. Actually, if an integral is saturated by distances of $1/\lambda$ order, then at $p \sim \lambda$, the integrand denominator has at least one power of momentum more than the product of the numerator and the measure of integration. In other words, any "long-distance" contribution (determined by the scale of $\lambda$) {\em has} to contain a positive power of the photon mass in its denominator. \section{Checking of the Results} \subsection{Long-Distance Contribution} To be certain that the long-distance contributions are found correctly, all of them were rederived in the framework of the nonrelativistic approach which exploits the Schr\"odinger equation as a starting point. For the states with nonzero angular momenta we used the following procedure. All the contributions prove to have the same analytic structure in \mbox{$\kappa$}, namely \[ \mbox{$\Delta E$} = \fr{m^2\mbox{$\alpha$}^6}{Mn^3} \Sigma, \] where \begin{eqnarray} \Sigma &=& \fr{1}{|\mbox{$\kappa$}|^3} \left ( a_{\infty} + \fr{a_{1/2}}{\mbox{$\kappa$}-1/2} + \fr{a_{-1/2}}{\mbox{$\kappa$}+1/2} + \fr{a'_{-1/2}}{(\mbox{$\kappa$}+1/2)^2} + \fr{a_{-1}}{\mbox{$\kappa$}+1} + \fr{a_{-3/2}}{\mbox{$\kappa$}+3/2} \right ) \nonumber \\ && +\fr{1}{n\mbox{$\kappa$}^2} \left ( b_{\infty} + \fr{b_{-1/2}}{\mbox{$\kappa$}+1/2} \right ) \nonumber \\ && + \fr{1}{n^2|\mbox{$\kappa$}|} \left ( c_{\infty} + \fr{c_{1/2}}{\mbox{$\kappa$}-1/2} + \fr{c_{-1/2}}{\mbox{$\kappa$}+1/2} + \fr{c_{-3/2}}{\mbox{$\kappa$}+3/2} \right )+\fr{d}{n^3}, \end{eqnarray} Constants $a,b,c,d$ evaluated in the nonrelativistic approach for individual contributions as their asymptotic values at $\mbox{$\kappa$} \mbox{$\rightarrow$} \infty$ or residues at corresponding poles were then compared with the respective results obtained in the relativistic approach. In the process of comparison, a number of 'nonrelativistic' contributions breaks down into groups according to the meaning of respective 'relativistic' ones. For example, the retardation part of the magnetic contribution (\ref{ret}) comprises three terms in the nonrelativistic approach: $\mbox{$\Delta E$}_{MC}^{(1)}$, $\mbox{$\Delta E$}_{MCC}^{(1)}$ and $\mbox{$\Delta E$}_{ret}^{(1)}$ (notations are from ref.\cite{p}). As we mentioned earlier, $S$ states should be treated separately in order to avoid fictious contributions arising due to the compensation between vanishing angular averages and linearly divergent radial ones. All the ultraviolet divergencies in $S$ states are checked to cancel each other. To this end we regularize the effective potentials which are too singular at $r \mbox{$\rightarrow$} 0$ and ensure that the total long-distance contribution for $S$ states is independent of the regularization parameter. \subsection{Short-Distance Contribution} In order to compare the results of the present work with those of \cite{PG}, the short-distance contribution was calculated using the Coulomb gauge also. A mass of the magnetic quantum was used as the infrared regulator. The scheme of calculation is completely analogous to those used previously in the case of the Feynman gauge. The short-distance contributions are: \begin{equation}\label{Csd} C_G = C_{\psi} = \fr{\pi\mbox{$\alpha$}^3\psi^2}{Mm} \fr{1}{2}; \end{equation} \medskip \begin{eqnarray} M_G &=& - \fr{\pi\mbox{$\alpha$}^3\psi^2}{Mm} \left ( \ln\fr{\lambda}{2m}+\fr{3}{2} \right ), \\ M_{\psi} &=& - \fr{\pi\mbox{$\alpha$}^3\psi^2}{Mm} \ln\fr{\lambda}{2m}; \end{eqnarray} \medskip \begin{eqnarray} S_G &=& \fr{\pi\mbox{$\alpha$}^3\psi^2}{Mm} \left ( 4\ln 2 - 2 \right ), \\ S_{\psi} &=& \fr{\pi\mbox{$\alpha$}^3\psi^2}{Mm} 2 \ln\fr{\lambda}{2m}.\label{Ssd} \end{eqnarray} Here $C$, $M$, $S$ denote Coulomb, magnetic and seagull contributions respectively. It is easy to check that the sum of these contributions coincides with (\ref{sd}). \section{Conclusion} Numerically, the correction to the energy equals 2.77 kHz for the ground state and 0.51 kHz for the $2S$ state. Being somewhat less than the na\"{\i}ve estimate ($m^2\mbox{$\alpha$}^6/M \approx 10.2$ kHz) it nevertheless is quite comparable with the accuracy of the near future measurements. It is also interesting to note that the corrections to $2S$ and $2P$ levels (with the radiative-recoil correction to $2P$ level \cite{p} taken into account) are rather close to each other, so that the correction to their difference, 0.04 kHz, can be considered as negligibly small at the present level of the experimental accuracy \cite{exp}. Let us now set up a correspondence between the results of the present work and those of the other papers. The result for $l>0$ levels appears to be firmly established \cite{p,ASY}. For $S$ levels our result is contradictory to the recent results of the analytic \cite{PG} and numerical \cite{ASY} calculations. To elucidate the origin of the disagreement, we consider the correction to the ground state energy. It is easy to verify that our short-distance results (\ref{Csd})--(\ref{Ssd}) are in one-to-one correspondence with respective "high-energy" contributions from \cite{PG}. A similar statement is true for long-distance contributions (low- and intermediate-energy ones in notations of \cite{PG}), with one exception. The coefficient --2 from Eq.(68) of Ref.\cite{PG} for the intermediate-energy contribution to the retarded exchange by the magnetic quantum, differs from our result, --1 (in the same units $m^2\mbox{$\alpha$}^6/M$), which arises after trivial averaging in (\ref{ret}) over the ground state. Unfortunately, we have not managed to reproduce the coefficient --2 starting from Eq.(67) of Ref.(\ref{ret}). Furthermore, several arguments can be brought forward, that the result (68),\cite{PG} for the retardation contribution looks at least suspicious. In particular, the logarithmic divergency in the order $m^2\mbox{$\alpha$}^6/M$ is known to appear due to the relativistic corrections to the {\em instant} transverse exchange. The result of the present work concerning the origin of this logarithmic divergency and the value of the corresponding coefficient is contained in (\ref{M-}) and is in complete agreement with those of Refs.\cite{PhScr} and \cite{DGE}. As for the effect of retardation, it gives rise to the finite contribution only (in accord with (\ref{ret})). But it follows from the result of Ref.\cite{PG} that just the retardation is the source of not only logarithmic, but even the linear divergency at small distances, while the long-distance relativistic correction to the instant transverse exchange does not contribute at all. \bigskip {\bf Acknowledgments}\nopagebreak The author is grateful to A. Vainshtein for stimulating discussions, and to M. Eides, S. Karshenboim, and V. Shabaev for their interest to the work. The work was supported by a fellowship of INTAS Grant 93-2492 and is carried out within the research program of International Center for Fundamental Physics in Moscow. Partial support from the program ``Universities of Russia'', Grant 94-6.7-2053, is also gratefully acknowledged. \newpage
\section{Introduction} In this article, we study homotopy groups and some related problems by using simplicial homotopy theory. The point of view here is that combinatorial aspects of group theory provide further information about homotopy groups.\\ \par The homotopy groups of the $3$-sphere, the suspension of $K(\pi,1)$ and wedges of $2$-spheres are shown to be the centers of certain groups with specific generators and specific relations. We list two group theoretical descriptions of $\pi_*(S^3)$ as follows. \begin{definition} {\bf A bracket arrangement of weight $n$} in a group is a set of elements defined recursively as follows: \par Let $G$ be a group and let $a_1,a_2,\cdots,a_n$ be a finite sequence of elements of $G$. Let $$ \beta^1(a_1)=a_1 $$ and if $n>1$, then let $$ \beta^n(a_1,\cdots,a_n)=[\beta^k(a_1,\cdots,a_k),\beta^l(a_{k+1},\cdots,a_n)] $$ for some $k$ and $l>0$. \end{definition} \begin{theorem} For $n\geq 1$, $\pi_{n+2}(S^{3})$ is isomorphic to the center of the group with generators $y_{0},\dots,y_{n}$ and relations $$[y_{i_{1}}^{\epsilon_{1}},y_{i_{2}}^{\epsilon_{2}},\dots,y_{i_{t}}^{\epsilon_{t}}]$$ with $\{i_{1},\dots,i_{t}\}=\{-1,0,\dots,n\}$ as sets in which the indices $i_{j}$ can be repeated, where $\epsilon_{j}=\pm 1, y_{-1}=(y_{0}\dots y_{n})^{-1}$ and the commutator bracket $[ \dots ]$ runs over all bracket arrangements of weight $t$ for each $t$. \end{theorem} \begin{notations} Let $G$ be a group and let $S$ be a subset of $G$. Let $<\!S\!>$ denote the normal subgroup generated by $S$. Let $H_{j}$ be a sequence of subgroups of $G$ for $1\leq j\leq k$. Let $[[H_{1},\dots,H_{k}]]$ denote the subgroup of $G$ generated by all of the commutators $[h_{i_1}^{(1)},\dots,h_{i_t}^{(t)}]$ with $\{i_1,\dots,i_t\}=\{1,\dots,k\}$ as sets in which the indices $i_j$can be repeated, where $h_j^{(s)}\in H_j$ and the commutator bracket $[ \dots ]$ runs over all of the bracket arrangements of weight $t$ for each $t$. \end{notations} \begin{theorem} Let $n\geq 1$. In the free group $F(y_0,\dots,y_n)$ freely generated by $y_0,\allowbreak \dots,\allowbreak y_n$ we have $$ ([[<\!y_0\!>,<\!y_1\!>,\dots,<\!y_n\!>]]\cap <\!y_{-1}\!>)/ [[<\!y_{-1}\!>,<\!y_0\!>,\dots,<\!y_n\!>]]\cong \pi_{n+2}(S^{3}) $$ where $y_{-1}=(y_0\dots y_n)^{-1}$. \end{theorem} The method of the proofs of these theorems is to study the Moore chain complex of Milnor's construction $F(S^1)$ for the $1$-sphere $S^1$. A group theoretical description of the homotopy groups $\pi_*(\Sigma K(\pi,1))$ is as follows. \begin{theorem} Let $\pi$ be any group and let $\{x^{(\alpha)}|\alpha\in J\}$ be a set of generators for $\pi$. Then, for $n\not=1$, $\pi_{n+2}(\Sigma K(\pi,1))$ is isomorphic to the center of the quotient group of the free product $${\coprod^{groups}}_{0\leq j\leq n}(\pi)_j$$ modulo the relations $$ [y^{(\alpha_1)\epsilon_1}_{i_1},y^{(\alpha_2)\epsilon_2}_{i_2},\cdots,y^{(\alpha_t)\epsilon_t}_{i_t}] $$ with $\{i_1,i_2,\cdots,i_t\}=\{-1,0,1,\cdots,n\}$ as sets, where $(\pi)_j$ is a copy of $\pi$ with generators $\{x^{(\alpha)}_j|\alpha\in J\}$, $\epsilon_j=\pm1$, $y^{(\alpha)}_{-1}={x^{(\alpha)}_0}^{-1}$, $y_j=x^{(\alpha)}_j{x^{(\alpha)}_{j+1}}^{-1}$ for $1\leq j\leq n-1$, $y^{(\alpha)}_n=x^{(\alpha)}_m$ and the commutator bracket $[\cdots]$ runs over all of the commutator bracket arrangment of weight $t$ for each $t$. \end{theorem} In particular, for $\pi={\bf Z}/m$, we have \begin{corollary} $\pi_{n+2}(\Sigma K({\bf Z}/m,1))$ is isomorphic to the center of the group with generators $x_{0},\dots,x_{n}$ and relations: (1). $x_j^m$ for $0\leq j\leq n$ and (2). $$[y_{i_{1}}^{\epsilon_{1}},y_{i_{2}}^{\epsilon_{2}},\dots,y_{i_{t}}^{\epsilon_{t}}]$$ with $\{i_{1},\dots,i_{t}\}=\{-1,0,\dots,n\}$ as sets in which the indices $i_{j}$ can be repeated, where $\epsilon_{j}=\pm 1$, $y_{-1}=x_0^{-1}$, $y_j=x_{j-1}x_j^{-1}$ for $-1\leq j\leq n-1$, $y_n=x_n$ and the commutator bracket $[ \dots ]$ runs over all bracket arrangements of weight $t$ for each $t$. \end{corollary} One of the features of Theorems 1.2 and 1.5 is that homotopy groups embed in certain 'enveloping groups'. These 'enveloping groups' have systematic and uniform structure. The centers of these groups are of course more complicated to analyze. Theorem 1.4 is very similar to the combinatorial decription of J. H. C. Whitehead's conjecture [see, Bo2, pp.317]. These decriptions give a combinatorial question how to give a computable way to understand the quotient groups $R\cap S/[R,S]$ for certain subgroups $R$ and $S$ of a finite generated free group. Few informations about this question are known [see, Bo1]. The article is organized as follows.\\ \par In Section 2, we study some general properties of simplicial groups. Central extensions in the Moore-Postnikov systems will be considered. In Section 3, we study the intersection of certain subgroups in free groups. The proofs of Theorems 1.2 and 1.4 is given in Section 4, where Theorem 4.5 is Theorem 1.2 and Theorem 4.10 is Theorem 1.4. The proof of Theorem 1.5 is given in Section 5, where Theorem 5.9 is Theorem 1.5. In Section 6, we give some applications of our descriptions. One example is to compute the homotopy groups of the Cohen's construction on the $1$-sphere. A direct corollary is to give a short proof of the Milnor's counter example for the minimal simplicial groups.\\ \par The author would like to thank F. Cohen, B. Gray, J. Harper, D. Kan, M. Mahowald, J. Moore, P. May, S. Priddy and many other mathematicians for their kindly encouragement and helpful suggestions. The author is indebted to helpful discussions with J. C. Moore and F. R. Cohen. \section{Central Extensions in simplicial group theory} In this section, we study some general properties of simplicial groups. A simplicial set $K$ is called a Kan complex if it satisfies the ${\it extension\, condition}$,i.e, for any simplicial map $f:\Lambda^k[n]\to K$ has an extension $g:\Delta[n]\to K$, where $\Delta[n]$ is the standard $n$ simplex, $\Lambda^k[n]$ is the subcomplex of $\Delta[n]$ generated by all $d_i(\sigma_n)$ for $i\not=k$ and $\sigma_n$ is the nondegenerate $n$ simplex in $\Delta[n]$,[K1, C2]. Recall that any simplicial group is a Kan complex [Mo1]. Given a simplicial group $G$, the Moore chain complex $(NG,d_0)$ is defined by $NG_n=\bigcap_{j\not=0}Kerd_j$ with $d_0:NG_n\to NG_{n-1}$. The classical Moore Theorem is that $\pi_n(G)\cong H_n(NG)$ [Mo1, Theorem 4;C2, Theorem 3.7 or K2, Proposition 5.4]. Now let ${\cal Z}_n={\cal Z}_n(G)=\bigcap_jKerd_j$ denote the cycles and let ${\cal B}_n={\cal B}_n(G)=d_0(NG_{n+1})$ denote the boundaries. It is easy to check that ${\cal B}_n$ is a normal subgroup of $G_n$ for any simplicial group $G$. \begin{lemma} Let $G$ be a simplicial group. Then the homotopy group $\pi_n(G)$ is contained in the center of $G_n/{\cal B}G_n$ for $n\geq1$. \end{lemma} \noindent{\em Proof.} Notice that $\pi_n(G)\cong{\cal Z}G_n/{\cal B}G_n$. It suffices to show that the commutator $[x,y]\in{\cal B}G_n$ for any $x\in{\cal Z}G_n$ and $y\in G_n$. \par Notice that $x$ is a cycle. There is a simplicial map $f_x: S^n\rightarrow G$ such that $f_x(\sigma_n)=x$, where $S^n$ is the standard $n$-sphere with a nondegenerate $n$-simplex $\sigma_n$. Now let the simplicial map $f_y:\Delta[n]\rightarrow G$ be the representative of $y$, i.e, $f_y(\tau_n)=y$ for the nondegenerate $n$-simplex $\tau_n$. Let $\phi$ be the composite $$ S^n\stackrel{j}{\rightarrow} S^n\wedge\Delta[n]\stackrel{[f_x,f_y]}{\rightarrow}G, $$ where $j(\sigma_n)=\sigma_n\wedge\tau_n$ and $[f_x,f_y](a\wedge b)=[f_x(a),f_y(b)]$, the commutator of $f_x(a)$ and $f_y(b)$. Notice that $\phi(\sigma_n)=[x,y]$ and $S^n\wedge\Delta[n]$ is contactible. Thus $[x,y]\in{\cal B}_n$. The assertion follows. \begin{definitions} Let $G$ be a simplicial group. The subsimplicial group {\bf $R_nG$} is defined by setting $$ (R_nG)_q=\{x\in G_q|f_x(\Delta[q])^{[n]}=1\}, $$ where $f_x$ is the representative of $x$ and $X^{[n]}$ is the $n$-skeleton of the simplicial set $X$. The subsimplicial group {\bf $\overline{R}_nG$} is defined by setting $$ (\overline{R}_nG)_q=\{x\in G_q|f_x((\Delta[q])_n)\subseteq{\cal B}G_n\}. $$ Let {\bf $P_nG$} denote $G/R_nG$ and let {\bf $\overline{P}_nG$} denote $G/\overline{R}_nG$. \end{definitions} It is easy to check that both $R_nG$ and $\overline{R}_nG$ are normal subsimplicial group of $G$. Thus both $P_nG$ and $\overline{P}_nG$ are quotient simplicial group of $G$. Notice that $$ R_nG\subseteq\overline{R}_nG\subseteq R_{n-1}G. $$ The quotient simplicial group $\overline{P}_nG$ is between $P_nG$ and $P_{n-1}G$. By checking the definition of Moore-Post\-nikov systems of a simplicial set [Mo1;C2], we have \begin{lemma} The quotient simplicial group $P_nG$ is the standard $n$-th Moore-Postnikov system of the simplicial group $G$. \end{lemma} The quotient simplicial group $\overline{P}_nG$ has the same homotopy type of $P_nG$. \begin{proposition} The quotient simplicial homomorphism $q_n:P_nG\rightarrow\overline{P}_nG$ is a homotopy equivalence for each $n$. \end{proposition} \noindent{\em Proof:} The Moore chain complex of $P_nG$ is as follows: $$ N(P_nG)_q=\cases{1&for\,$q>n+1$,\cr NG_{n+1}/{\cal Z}G_{n+1}&for\,$q=n+1$,\cr NG_q&for\,$q<n+1$.\cr} $$ The Moore chain complex of $\overline{P}_nG$ is as follows: $$ N(\overline{P}_nG)_q=\cases{1&for\,$q>n$,\cr NG_n/{\cal B}G_n &for\, $q=n$,\cr NG_q &for $q<n$.\cr} $$ Thus $(q_n)_*:\pi_*(P_nG)\rightarrow\pi_*(\overline{P}_nG)$ is an isomorphism and the assertion follows.\\ \par Let $\overline{F}_n$ denote the kernel of the quotient simplicial homomorphism $r_n:\overline{P}_nG\rightarrow P_{n-1}G$ for $n>0$. \begin{proposition} The simplicial group $\overline{F}_n$ is the minimal simplicial group $K(\pi_nG,n)$ for each $n>0$. \end{proposition} \noindent{\em Proof:} It is directly to check that $$ N(\overline{F}_nG)_q=\cases{1&for\,$q\not=n$,\cr \pi_nG&for\, $q=n$.\cr} $$ The assertion follows. \begin{proposition} The short exact sequence of simplicial groups $$ 0\rightarrow\overline{F}_nG\rightarrow\overline{P}_nG\rightarrow P_{n-1}G\rightarrow0 $$ is a central extension for each $n>0$. \end{proposition} \noindent{\em Proof:} Consider the relative commutator sussimplicial group $[\overline{F}_nG,\overline{P}_nG]$. By Lemma 2.1, $[\overline{F}_nG,\overline{P}_nG]_n=1$. Notice that $[\overline{F}_nG,\overline{P}_nG]$ is a subsimplicial group of minimal simplicial group $\overline{F}_nG\cong K(\pi_nG,n)$. Thus $[\overline{F}_nG,\overline{P}_nG]=1$ and the assertion follows. \begin{definition} A simplicial group is said to be {\it $r$-centerless} if the center $Z(G_n)=\{1\}$ for $n\geq r$. \end{definition} \begin{proposition} Let $G$ be a reduced $r$-centerless simplicial group. Then $\pi_n(G)\cong Z(G_n/{\cal B}_n)$ for $n\geq r+1$ \end{proposition} \noindent{\em Proof.} By Lemma 2.1, ${\cal Z}_n/{\cal B}_n\subseteq Z(G_n/{\cal B}_n)$. It suffices to show that $Z(G_n/{\cal B}_n)\subseteq{\cal Z}_n/{\cal B}_n$ for $n\geq r+1$. Now let ${\tilde{x}}\in Z(G_n/{\cal B}_n)$. Choose $x\in G_n$ with $p(x)={\tilde x}$, where $p:G_n\to G_n/{\cal B}_n$ is the quotient homomorphism. To check ${\tilde x}\in{\cal Z}_n/{\cal B}_n$, it suffices to show that $x\in {\cal Z}_n$ or $d_jx=1$ for all $j$. Since $Z(G_{n-1})=\{1\}$, $d_jx=1$ if and only if $[d_jx,y]=1$ for all $y\in G_{n-1}$. Now $[d_jx,y]=d_j[x,s_{j-1}y]$ for $j>0$ and $[d_0x,y]=d_0[x,s_0y]$. Since ${\tilde x}\in Z(G_n/{\cal B}_n)$, $[x,z]\in{\cal B}_n\subseteq{\cal Z}_n$ for all $z\in G_n$ and therefore $[d_jx,y]=1$ for all $y\in G_{n-1}$. The assertion follows.\\ \par By inspecting the proof, we also have \begin{proposition} Let $G$ be a reduced $r$-centerless simplicial group. Then $$Z(G_n/{\cal Z}_n)=\{1\}$$ for $n\geq r+1$. \end{proposition} \begin{lemma} Let $G$ be a reduced simplicial group so that $G_n$ is cyclic or centerless for each n. Then there exists a unique integer $\gamma_G>0$ so that $G_n=\{1\}$ for $n<\gamma_G$ and $Z(G_n)=\{1\}$ for $n>\gamma_G$. \end{lemma} \noindent{\em Proof.} Let $\gamma_G=max\{\gamma|G_n=\{1\}\,\,for\,\,n<\gamma\}$. Then $\gamma_G>0$. If $\gamma_G<\infty$. Then $G_{\gamma_G}\neq\{1\}$. We show that $G_{\gamma_G+q}$ is centerless for each $q>0$. Notice that $d_0^q\circ s_0^q:G_n\rightarrow G_{n+q}\rightarrow G_n$ and $d_1^q\circ s_1^q:G_n\rightarrow G_{n+q}\rightarrow G_n$ are identities for $n>0$ Thus $s_0^q(G_{\gamma_G})$ and $s_1^q(G_{\gamma_G})$ are nontrivial summands of $G_{\gamma_G+q}$. Now let $x=s_0^qy=s_1^qz\in s_0^q(G_{\gamma_G})\cap s_1^q(G_{\gamma_G})$. Then $d_{q+1}x=d_{q+1}s_0^qy=s_0^qd_1y=1=d_{q+1}s_1^qz=s_1^{q-1}d_2s_1z=s_1^{q-1}z$. Thus $x=s_1^qz=1$. And therefore $s_0^q(G_{\gamma_G})\cap s_1^q(G_{\gamma_G})=\{1\}$. The assertion follows. \begin{corollary} Let $G$ be a reduced simplicial group such that $G_n$ is cyclic or centerless for each $n$. Then $\pi_n(G)\cong Z(G_n/{\cal B}_n)$ for $n\neq\gamma_G+1$, where $\gamma_G$ is defined as above. \end{corollary} Notice that, for any free group $F$, $rank(F)\geq2\Leftrightarrow Z(F)=\{1\}$ and $F\neq\{1\}$. We have \begin{lemma} Let $G$ be a reduced simplicial group such that $G_n$ is a free group for each $n$. Then there exits a unique integer $\gamma_G>0$ so that $G_n=\{1\}$ for $n<\gamma_G$ and $rank(G_n)\geq2$ for $n>\gamma_G$. \end{lemma} \begin{example} The $1$-stem is determined in this example. \end{example} Let $G=F(S^n)$, Milnor's $F$-construction on the standard n-sphere for $n\geq1$. Then $G_n\cong F(\sigma)\cong {\bf Z}(\sigma)$, the free abelian group generated by $\sigma$, $$G_{n+1}\cong F(s_0\sigma,s_1\sigma,\dots,s_n\sigma)$$ and $G_{n+2}\cong F(s_is_j\sigma |0\leq j<i\leq n)$. It is easy to check that $\Gamma^2G_{n+1}={\cal Z}_{n+1}$, where $\Gamma^qG$ is the $q$-th term in the lower central series of a group $G$ starting with $\Gamma^1G=G$. By Lemma 2.1, $$\Gamma^3G_{n+1}=[{\cal Z}_{n+1}, G_{n+1}]\subseteq{\cal B}_{n+1}.$$ If $n=1$, then it will be shown that $\Gamma^3G_{n+1}={\cal B}_{n+1}$ in Section 4 and therefore $\pi_3(S^2)\cong\pi_2(F(S^1))\cong{\bf Z}$, which is generated by $[s_0\sigma,s_1\sigma]$. \par Suppose that $n>1$. Consider the following equations $$ d_k([s_{j-1}s_i\sigma,s_{j+1}s_j\sigma])=\cases{1&for\,$k\neq j,$\cr[s_i\sigma,s_j\sigma]&for\,k=j,\cr} $$ for $i+1<j\leq n$, $$ d_k[s_{i+2}s_{i+1}\sigma,s_{i+3}s_i\sigma]=\cases{[s_{i+1}\sigma,s_{i+2}\sigma]&for\,k=i+1,\cr[s_{i+1}\sigma,s_i\sigma]&for\,k=i+3,\cr1&otherwise,\cr} $$ and $$ d_k[s_{i+2}s_i\sigma,s_{i+3}s_{i+1}\sigma]=\cases{[s_{i+1}\sigma,s_{i+2}\sigma]&for\,k=i+1,\cr[s_i\sigma,s_{i+2}\sigma]&for\,k=i+2,\cr[s_i\sigma,s_{i+1}\sigma]&for\,k=i+3,\cr1&otherwise.\cr} $$ By the Homotopy Addition Theorem [C2, Theorem 2.4], $[s_i\sigma, s_j\sigma]\in{\cal B}_{n+1}$ for $i+1<j$, $[s_{i+1}\sigma,s_{i+2}\sigma]\equiv[s_i\sigma,s_{i+1}\sigma]$ mod ${\cal B}_{n+1}$ and $0\equiv[s_i\sigma,s_{i+2}\sigma]\equiv[s_{i+1}\sigma,s_{i+2}\sigma]\equiv 2[s_i\sigma,s_{i+1}\sigma]$ if $i+2\leq n$. Notice that $[s_0\sigma,s_1\sigma]\neq0$ in $\pi_{n+1}(\Gamma^2G/\Gamma^3G)$. Thus $[s_0\sigma,s_1\sigma]\notin{\cal B}_{n+1}$ and, by the relations above, $$ \pi_{n+2}(S^{n+1})\cong\pi_{n+1}(G)\cong{\cal Z}_{n+1}/{\cal B}_{n+1}\cong{\bf Z}/2 $$ for $n\geq2$, which can be represented by $[s_i\sigma,s_{i+1}\sigma]$ for $0\leq i\leq n-1$. \section{Intersections of certain subgroups in free groups} In this section, we give some group theory preliminary. The intersections of certain subgroups in free groups are considered in this section. We will use these informations to determine the Moore chain complex of certain simplicial groups in the next sections. \begin{definition} let $S$ be a set and let $T\subseteq S$ a subset. The {\bf projection homomorphism} $$\pi:F(S)\rightarrow F(T)$$ is defined by $$ \pi(x)=\cases{x&$x\in T$,\cr 1&$x\notin T$.\cr} $$ \end{definition} Now let $\pi:F(S)\rightarrow F(T)$ be a projection homomorphism and let $R$ equal the kernel of $\pi$. Define the subsets of the free group $F(S)$ as follows. $$ {\cal A}_T(k)=\{[[x,y_1^{\epsilon_1}]\cdots],y_t^{\epsilon_t}]|0\leq t\leq k,\epsilon_j=\pm1, y=y_1^{\epsilon_1}\dots y_t^{\epsilon_t}\in F(T),x\in S-T\}, $$ where $y=y_1^{\epsilon_1}\cdots y_t^{\epsilon_t}\in F(T)$ runs over reduced words in $F(T)$ with $t\leq k$ and $y_j\in T$. Furthermore define $[[x,y_1^{\epsilon_1}]\dots],y_t^{\epsilon_t}]=x$ for $t=0$. Define $$ {\cal B}_T(k)=\{\phi^{-1}x\phi|\phi\in F(T)\,a\,reduced\,word\,with\,lenth\,l(\phi)\leq k,x\in S-T\}, $$ $${\cal A}_T=\cup_{k\geq0}{\cal A}_T(k)$$ and $${\cal B}_T=\cup_{k\geq0}{\cal B}_T(k).$$ By the classical Kurosch-Schreier theorem ( see [MKS, pp.243, K2, Theorem 18.1]), we have \begin{proposition} The subgroup $R$ is a free group freely generated by ${\cal B}_T$. \end{proposition} We will show that ${\cal A}_T$ is also a set of free generators for $R$. We need a lemma. \begin{lemma} Let $\phi:F_1\rightarrow F_2$ be a homomorphism of free groups. Suppose that $\phi^{ab}:F^{ab}_1\rightarrow F^{ab}_2$ is an isomorphism, where $F^{ab}$ is the abelianlizer of the group $F$. Then $\phi:F_1\rightarrow F_2$ is a monomorphism. \end{lemma} \noindent{\em Proof.} Notice that $\phi_*: H_*(F_1)\rightarrow H_*(F_2)$ is an isomorphism, where $H_*(G)$ is the homology of the group $G$. Thus $F_1/\Gamma^rF_1\rightarrow F_2/\Gamma^r F_2$ is an isomorphism for each $r$, where $\Gamma^rG$ is the $r$-th term in the lower central seris of the group $G$ starting with $\Gamma^1G=G$ and so $$ lim_r F_1/\Gamma^r F_1\rightarrow lim_r F_2/\Gamma^r F_2$$ is an isomorphism. Notice that $\cap_r\Gamma^rF=1$ for any free group $F$. Thus $F\rightarrow lim_r F/\Gamma^r F$ is a monomorphism. The assertion follows.\\ \par \begin{proposition} The subgroup $R$ is a free group freely generated by ${\cal A}_T$. \end{proposition} \noindent{\em Proof.} First we assume that both $S$ and $T$ are finite sets. Denote by $i_k:{\cal A}_T(k)\rightarrow R$ and $j_k:{\cal B}_T(k)\rightarrow R$ the natural inclusions. Notice that $$R=F({\cal B}_T)=colim_kF({\cal B}_T(k)).$$ We set up the following steps.\\ \par \noindent{\em Step1.} ${\cal A}_T(k)\subseteq F({\cal B}_T(k))$.\\ \par The proof of this statement is given by induction on $k$ starting with ${\cal A}_T(0)={\cal B}_T(0)=S-T$. Suppose that ${\cal A}_T(k-1)\subseteq F({\cal B}_T(k-1))$ and let $w=[[x,y_1^{\epsilon_1},\cdots,y_t^{\epsilon_t}]\in{\cal A}_T(k)$. If $t<k$, then $w\in{\cal A}_T(k-1)\subseteq F({\cal B}_T(k-1))\subseteq F({\cal B}_T(k))$, by induction. Now $$ [[x,y_1^{\epsilon_1}]\cdots],y_k^{\epsilon_k}]= [[x,y_1^{\epsilon_1}]\cdots],y_{k-1}^{\epsilon_{k-1}}]^{-1}\cdot y_k^{-\epsilon_k}[[x,y_1^{\epsilon_1}]\cdots],y_{k-1}^{\epsilon_{k-1}}]y_k^{\epsilon_k} $$ Now since $[[x,y_1^{\epsilon_1}]\cdots],y_{k-1}^{\epsilon_{k-1}}]\in F({\cal B}_T(k-1))$, $ [[x,y_1^{\epsilon_1}]\cdots],y_{k-1}^{\epsilon_{k-1}}]=\prod_{j=1}^s(\phi_j^{-1}x_j\phi_j))^{\eta_j} $ with $\phi_j^{-1}x_j\phi_j\in{\cal B}_T(k-1)$ and $\eta_j=\pm1$. Thus $$ w=(\prod_{j=1}^s(\phi_j^{-1}x_j\phi_j))^{\eta_j})^{-1}\cdot\prod_{j=1}^s(y_k^{-\epsilon_k}\phi_j^{-1}x_j\phi_j y_k^{\epsilon_k}))^{\eta_j}\in F({\cal B}_T(k)) $$ The induction is finished.\\ \par \noindent{\em Step 2.} ${\tilde i_k}: F({\cal A}_T(k))\rightarrow F({\cal B}_T(k))$ is an epimorphism, where the homomorphism ${\tilde i_k}$ is induced by the inclusion $i_k:{\cal A}_T(k)\rightarrow F({\cal B}_T(k))$\\ \par The proof of this step is given induction on $k$ starting with $F({\cal A}_T(0))=F({\cal B}_T(0))=F(S-T)$. Suppose that $F({\cal A}_T(k-1))\rightarrow F({\cal B}_T(k-1))$ is an epimorphism and consider ${\tilde i_k}:F({\cal A}_T(k))\rightarrow F({\cal B}_T(k))$. Let $\phi^{-1}x\phi\in{\cal B}_T(k)$, where $\phi=y_1^{\epsilon_1}\cdots y_t^{\epsilon_t}$ is a reduced word with $t\leq k$. If $t\leq k-1$, then $\phi^{-1}x\phi\in Im\varphi_k$ by induction. Let $\phi=y_1^{\epsilon_1}\cdots y_k^{\epsilon_k}$ be a reduced word and let $z$ denote the word $(y_1^{\epsilon_1}\cdots y_{k-1}^{\epsilon_{k-1}})^{-1}xy_1^{\epsilon_1}\cdots y_{k-1}^{\epsilon_{k-1}}$. Then $\phi^{-1}x\phi=z\cdot[z,y_k^{\epsilon_k}]$. Notice that $z\in Im({\tilde i_k})$ by induction. It suffices to show that $[w,y^{\epsilon}]\in F({\cal A}_T(k))$ for $w\in{\cal A}_T(k-1)$ for all $w\in{\cal A}_T(k-1)$, $y\in T$ and $\epsilon=\pm1$ by the Witt-Hall identity that $$ [ab,c]=[a,c]\cdot[[a,c],b]\cdot[b,c]. $$ We show this by second induction starting with $$ {\cal A}_T(1)=\{[x,y^{\epsilon}]|y\in T, x\in S-T,\epsilon=\pm1\}. $$ Let $w=[[x,y_1^{\epsilon_1}],\cdots],y_t^{\epsilon_t}]$ be in ${\cal A}_T(k-1)$ with $k>1$, where $y^{\epsilon_1}_1\cdots y_t^{\epsilon_t}$ is a reduced word. Let $y\in T$ and let $\epsilon=\pm1$. If $y_1^{\epsilon_1}\cdots y_t^{\epsilon_t}y^{\epsilon}$ is a reduced word, then $[w,y^{\epsilon}]\in F({\cal A}_T(k))$ by definition. Suppose that $y_1^{\epsilon_1}\cdots y_t^{\epsilon_t}y^{\epsilon}$ is not a reduced word. Then $t>0$, $y=y_t$ and $\epsilon=-\epsilon_t$. Let $w'$ denote $[[x,y_1^{\epsilon_1}],\cdots,],y_{t-1}^{\epsilon_{t-1}}]\in{\cal A}_T(k-2)$. Then $w=[w',y_t^{\epsilon_t}]$. By the Witt-Hall identities, there is an equation $$ 1=[w',y_t^{-\epsilon_t}]\cdot w\cdot [w,y_t^{-\epsilon_t}]. $$ By induction, $[w',y_t^{-\epsilon_t}]\in F({\cal A}_T(k-1))\subseteq F({\cal A}_T(k))$ and so $$[w,y_t^{-\epsilon_t}]=w^{-1}[w',y_t^{-\epsilon_t}]^{-1}\in F({\cal A}_T(k-1))\subseteq F({\cal A}_T(k)).$$ The second induction is finished and so the first induction is finished. The assertion follows.\\ \par \noindent{\em Step 3.} ${\tilde i_k}: F({\cal A}_T(k))\rightarrow F({\cal B}_T(k))$ is an isomorphism.\\ \par By Step 2, ${\bf Z}({\cal A}_T(k))\rightarrow{\bf Z}({\cal B}_T(k))$ is an epimorphism. Notice that $$rank(F({\cal B}_T(k)))=|{\cal B}_T(k)|=|{\cal A}_T(k)|=rank(F({\cal A}_T(k))).$$ Thus ${\tilde i_k}:{\bf Z}({\cal A}_T(k))\rightarrow{\bf Z}({\cal B}_T(k))$ is an isomorphism and so ${\tilde i_k}: F({\cal A}_T(k))\rightarrow F({\cal B}_T(k))$ is a monomorphism. Thus ${\tilde i_k}$ is an isomorphism.\\ \par \noindent{\em Step 4.} Since $F({\cal A}_T(k))\rightarrow F({\cal B}_T(k))$ is an isomorphism for each $k$, $F({\cal A}_T)=colim_kF({\cal A}_T(k))\rightarrow F({\cal B}_T)=colim_kF({\cal B}_T(k))$ is an isomorphism.\\ \par Now consider the general case. By Lemma 3.6, it suffices to show that ${\tilde i}: F({\cal A}_T)\rightarrow F({\cal B}_T)$ is an isomorphism. To check that $F({\cal A}_T)\rightarrow F({\cal B}_T)$ is an epimorphism. Let $w\in{\cal B}_T$, there exist finite subsets $S'$ and $T'$ of $S$ and $T$, respectively, so that $w\in{\cal B}_{T'}$. By the special case as above, ${\tilde i}|_{F({\cal A}_{T'})}: F({\cal A}_{T'})\rightarrow F({\cal B}_{T'})$ is an isomorphism and $w\in Im{\tilde i}|_{F({\cal A}_{T'})}$. Thus ${\tilde i}$ is an epimorphism. To check that $F({\cal A}_T)\rightarrow F({\cal B}_T)$ is a monomorphism. Let $w\in Ker{\tilde i}$, there exist finite subsets $S'$ and $T'$ of $S$ and $T$, respectively, so that $w\in F({\cal A}_{T'})$. Notice that ${\tilde i}|_{F({\cal A}_{T'})}$ is an isomorphism. Thus $w=1$ and the assertion follows.\\ \par Now let's consider the intersection of kernels of projection homomorphisms. Let $S$ be a set and let $T_j$ be a subset of $S$ for $1\leq j\leq k$. Let $\pi_j:F(S)\rightarrow F(T_j)$ be the projection homomorphism for $1\leq j\leq k$. We construct a subset ${\cal A}(T_1,\cdots,T_k)$ of $F(S)$ by induction on $k$ as follows. $$ {\cal A}(T_1)={\cal A}_{T_1}, $$ where ${\cal A}_T$ is defined in Definition3.1. Let $$ T_2^{(2)}=\{w\in{\cal A}(T_1)|w=[[x,y_1^{\epsilon_1},\cdots,y_t^{\epsilon_t}]\, with\,x,y_j\in T_2\, for\,all\,j\} $$ and define $$ {\cal A}(T_1,T_2)={\cal A}(T_1)_{T_2^{(2)}}. $$ Suppose that ${\cal A}(T_1,T_2,\cdots,T_{k-1})$ is well defined so that all of the elements in ${\cal A}(T_1,T_2,\cdots,T_{k-1})$ are written down as certain commutators in $F(S)$ in terms of elements in $S$. Let $T_k^{(k)}$ be the subset of ${\cal A}(T_1,T_2,\cdots,T_{k-1})$ defined by $$ T^{(k)}_k=\{w\in{\cal A}(T_1,T_2,\cdots,T_{k-1})|w=[x_1^{\epsilon_1},\cdots,x_l^{\epsilon_l}]\,with\,x_j\in T_k\,for\,all\,j\}, $$ where $[x_1^{\epsilon_1},\cdots,x_l^{\epsilon_l}]$ are the elements in ${\cal A}(T_1,T_2,\cdots,T_{k-1})$ which are written down as commutators. Then define $$ {\cal A}(T_1,T_2,\cdots,T_k)={\cal A}(T_1,T_2,\cdots,T_{k-1})_{T_k^{(k)}} $$ \begin{theorem} Let $S$ be a set and let $T_j$ be a subset of $S$ for $1\leq j\leq k$. Let $\pi_j:F(S)\rightarrow F(T_j)$ be the projection homomorphism for $1\leq j\leq k$. Then the intersection $\bigcap_{j=1}^kKer\pi_j$ is a free group freely generated by ${\cal A}(T_1,T_2,\cdots,T_k)$. \end{theorem} \noindent{\em Proof.} The proof is given by induction on $k$. If $k=1$, the assertion follows from the above lemma. Suppose that $\bigcap_{j=1}^{k-1}Ker\pi_j=F({\cal A}(T_1,T_2,\cdots,T_{k-1}))$ and consider $\pi_k:F(S)\rightarrow F(T_k)$. Then $$\bigcap_{j=1}^kKer\pi_j=Ker({\bar\pi_k}:F({\cal A}(T_1,T_2,\cdots,T_{k-1}))\rightarrow F(T_k))$$, where ${\bar\pi_k}$ is $\pi_k$ restricted to the subgroup $F({\cal A}(T_1,T_2,\cdots,T_{k-1}))$. Let $$w=[x_1^{\epsilon_1},\cdots,x_l^{\epsilon_l}]\in{\cal A}(T_1,T_2,\cdots,T_{k-1}).$$ If $w\notin T_k^{(k)}$, then $x_j\notin T_k$ for some $j$ and ${\bar\pi_k}(w)=1$. Thus ${\bar\pi_k}$ factors through $F(T_k^{(k)})$, i.e, there is a homomorphism $j:F(T_k^{(k)})\rightarrow F(T_k)$ so that ${\bar\pi_k}=j\circ\pi$, where $\pi:F({\cal A}(T_1,T_2,\cdots,T_{k-1}))\rightarrow F(T_k^{(k)})$ is the projection homomorphism. We claim that $j:F(T_k^{(k)})\rightarrow F(T_k)$ is a monomorphism. Consider the commutative diagram $$ \hspace{0.0in} \begin{array}{cccccccccc} &F({\cal A}(T_1,T_2,\cdots,T_{k-1}))&\stackrel{{\bar\pi_k}}{\rightarrow}& F(T_k)& \hookrightarrow& F(S) \\ & \uparrow & & \uparrow j & & \uparrow \\ & F(T_k^{(k)}) &\stackrel{=}{\rightarrow} & F(T_k^{(k)}) & \hookrightarrow & F({\cal A}(T_1,T_2,\cdots,T_{k-1})), \end{array} $$ where $F(T_k^{(k)})\rightarrow F({\cal A}(T_1,T_2,\cdots,T_{k-1}))$ and $F({\cal A}(T_1,T_2,\cdots,T_{k-1}))\rightarrow F(S)$ are inclusions of subgroups. Thus $j:F(T_k^{(k)})\rightarrow F(T_k)$ is a monomorphism and $$ Ker{\bar\pi_k}=Ker(F({\cal A}(T_1,T_2,\cdots,T_{k-1}))\rightarrow F(T_k^{(k)})=F({\cal A}(T_1,T_2,\cdots,T_k)). $$ The assertion follows. \begin{corollary} Let $\pi_j$ be the projection homomorphisms as in Theorem 3.5. Then the intersection subgroup $\bigcap_{j=1}^kKer\pi_j$ equals the commutator subgroup $[[<T_1>,\cdots,<T_k>]]$ which is defined in Notations 1.3. \end{corollary} \section{On the Homotopy Groups of the $3$-sphere} In this section, we study the Moore chain complex of the Milnor's construction $F(S^1)$. The proofs Theorems 1.2 and 1.4 are given in this section, where Theorem 4.5 is Theorem 1.2 and Theorem 4.10 is Theorem 1.4. Recall that the simplicial $1$-sphere $S^1$ is a free simplicial set generated by a $1$-simplex $\sigma$. Thus $S^1_0=\{*\}$, $S^1_1=\{\sigma,*\}$ and $S^1_{n+1}=\{*,s_n\cdots s_{i+1}s_{i-1}\cdots s_0\sigma|0\leq j\leq n\}$. Let $x_i$ denote $s_n\cdots s_{i+1}s_{i-1}\cdots s_0\sigma$. Then $F(S^1)_{n+1}=F(x_0,x_1,\cdots, x_n)$ the free group freely generated by $x_0,\cdots,x_n$. Let $y_i$ denote $x_{i-1}x_i^{-1}$ for $-1\leq i\leq n$, where we put $x_{-1}=x_{n+1}=1$ in $F(S^1)_{n+1}=F(x_0,\cdots, x_n)$. By direct calculation, we have \begin{lemma} $F(S^1)_{n+1}=F(y_0,\cdots,y_n)$ with $$ d_jy_k=\cases{y_{k-1}&$j\leq k$,\cr 1&$j=k+1$,\cr y_k&$j>k+1$,\cr} $$ and $$ s_jy_k=\cases{y_{k+1}&$j\leq k$,\cr y_ky_{k+1}&$j=k+1$,and\cr y_k&$j>k+1$\cr} $$ for $0\leq j\leq n+1$, where $y_{-1}=(y_0\cdots y_{n-1})^{-1}$ in $F(S^1)_n$. \end{lemma} Now let $C_{n+1}$ denote the subgroup of $F(y_0,\cdots,y_n)$ generated by all of the commutators $[y_{i_1}^{\epsilon_1},\cdots,y_{i_t}^{\epsilon_t}]$ with $\{i_1,\cdots,i_t\}=\{0,1,\cdots,n\}$ as sets, i.e. each $j$ $(0\leq j\leq n)$ appears in the index set $\{i_1,\cdots, i_t\}$ at least one time, where $\epsilon_j=\pm1$ and the commutator $[y_{i_1}^{\epsilon_1},\cdots,y_{i_t}^{\epsilon_t}]$ runs over all of the commutator bracket arrangements of weight $t$ for $y_{i_1}^{\epsilon_1},\cdots,y_{i_t}^{\epsilon_t}$. \begin{lemma} The group $C_{n+1}$ is a subgroup of $NF(S^1)_{n+1}$,i.e $C_{n+1}\subseteq \cap_{j\neq 0}Ker(d_j)$. \end{lemma} \noindent{\em Proof.} Notice that $d_jy_{j-1}=1$ for $1\leq j\leq n+1$. Since $\{y_{i_1},\cdots, y_{i_t}\}=\{y_0,y_1,\cdots,y_n\}$, $d_j[y_{i_1}^{\epsilon_1},\cdots,y_{i_t}^{\epsilon_t}]=1$ for each $j>0$. The assertion follows.\\ \par \begin{theorem} $NF(S^1)_{n+1}=C_{n+1}$. \end{theorem} \noindent{\em Proof.} For $1\leq j\leq n+1$, let $S=\{y_0,y_1,\cdots,y_n\}$ and let $T_j=\{y_0,\cdots,{\hat y_j}\cdots,y_n\}$. By Lemma 4.1, there is a commutative diagram $$ \hspace{0.0in} \begin{array}{cccccc} & F(S) & \stackrel{\pi_j}{\rightarrow} & F(T_{j-1}) \\ & \downarrow d_j & & {\bar d_j}\downarrow \cong \\ & F(y_0,\cdots,y_{n-1}) &\stackrel{=}{\rightarrow} & F(y_0,\cdots,y_{n-1}),\\ \end{array} $$ where $$ {\bar d_j}(y_k)=\cases{y_k&$0\leq k\leq j-2,$\cr y_{k-1}&$j\leq k\leq n.$\cr} $$ Thus $Kerd_j=Ker\pi_j$ and $NF(S^1)_{n+1}=\bigcap_{j=1}^{n+1}Kerd_j=\bigcap_{j=1}^{n+1}Ker\pi_j$. By Theorem 3.5, $NF(S^1)_{n+1}=F({\cal A}(T_0,T_1,\cdots,T_n))$, where the notation ${\cal A}(T_0,T_1,\cdots,T_n)$ is given in Section 3. To check that $$ F({\cal A}(T_0,T_1,\cdots,T_n))\subseteq C_{n+1}, $$ it suffices to show that ${\cal A}(T_0,T_1,\cdots,T_n)\subseteq C_{n+1}$. This will follow from the following statement.\\ \par \noindent{\em Statement.} For each $0\leq j\leq n$ and $w=[y_{i_1}^{\epsilon_1},y_{i_2}^{\epsilon_2},\cdots,y_{i_t}^{\epsilon_t}]\in{\cal A}(T_0,T_1,\cdots,T_j)$, $\{y_0,y_1,\cdots,y_j\}\subseteq\{y_{i_1},y_{i_2},\cdots,y_{i_t}\}$.\\ \par We show this statement by induction on $j$. Note that $F(T_0)=F(y_1,\cdots,y_n)$. For $j=0$, $$ {\cal A}(T_0)=\{[[y_0,y_{i_1}^{\epsilon_1}],\cdots,],y_{i_t}^{\epsilon_t}]|i_s>0, y_{i_1}^{\epsilon_1}\cdots y_{i_t}^{\epsilon_t}\,a\,reduced\,word\, in\,F(T_0)\}. $$ Thus the assertion holds for $j=0$. Suppose that the assertion holds for $j-1$ with $j\leq n$. Notice that $T_j=\{y_0,\cdots,{\hat y_j},\cdots, y_n\}$. Thus $$ T_j^{(j)}=\{w\in{\cal A}(T_0,T_1,\cdots,T_{j-1})|w=[y_{i_1}^{\epsilon_1},\cdots,y_{i_t}^{\epsilon_t}]\, with\,y_j\notin\{y_{i_1},\cdots,y_{i_t}\}\}. $$ and so $y_j\in\{y_{i_1},\cdots,y_{i_t}\}$ for $w=[y_{i_1}^{\epsilon_1},\cdots,y_{i_t}^{\epsilon_t}]\in{\cal A}(T_0,T_1,\cdots,T_{j-1})-T_j^{(j)}$. Hence, by induction, $$ \{y_0,y_1,\cdots,y_j\}\subseteq\{y_{i_1},y_{i_2},\cdots,y_{i_t}\} $$ for $w=[y_{i_1}^{\epsilon_1},\cdots,y_{i_t}^{\epsilon_t}]\in{\cal A}(T_0,T_1,\cdots,T_{j-1})-T_j^{(j)}$. Notice that $${\cal A}(T_0,T_1,\cdots,T_j)={\cal A}(T_0,T_1,\cdots,T_{j-1})_{T_j^{(j)}}.$$ Thus $$ \{y_0,y_1,\cdots,y_j\}\subseteq\{y_{i_1},y_{i_2},\cdots,y_{i_t}\} $$ for each $w=[y_{i_1}^{\epsilon_1},\cdots,y_{i_t}^{\epsilon_t}]\in{\cal A}(T_0,T_1,\cdots,T_{j})$. The induction is finished and the theorem follows. \begin{corollary} ${\cal A}(T_0,T_1,\cdots,T_n)$ is a set of free generators for $NF(S^1)_{n+1}$. \end{corollary} \begin{theorem}[Theorem 1.2] For $n\geq1$, $\pi_{n+2}(S^3)$ is isomorphic to the center of the group with generators $y_0,y_1,\cdots,y_n$ and relations $[y_{i_1}^{\epsilon_1},y_{i_2}^{\epsilon_2},\cdots,y_{i_t}^{\epsilon_t}]$ with $$ \{i_1,i_2,\cdots,i_t\}=\{-1,0,1,\cdots,n\} $$ as sets, where $\epsilon_j=\pm1$,$y_{-1}=(y_0y_1\cdots y_n)^{-1}$ and the commutator bracket $[\cdots]$ runs over all bracket arrangements of weight $t$ for each $t$. \end{theorem} \noindent{\em Proof.} Notice that $\pi_{n+2}(S^3)\cong\pi_{n+1}F(S^1)$ for $n\geq1$ and ${\cal B}_{n+1}=d_0(NF(S^1)_{n+2})$. By Lemma 4.1 and Theorem 4.3, ${\cal B}_{n+1}$ is generated by $[y_{i_1}^{\epsilon_1},y_{i_2}^{\epsilon_2},\cdots,y_{i_t}^{\epsilon_t}]$ so that $\{y_{i_1},y_{i_2},\cdots,y_{i_t}\}=\{y_0,y_1,\cdots,y_n\}$, where $\epsilon_j=\pm1$. By Proposition 2.8, it suffices to check that $\pi_2(F(S^1))\cong Z(F(S^1)_2/{\cal B}_2)=Z(F(y_0,y_1)/{\cal B}_2)$. By Example 2.13, ${\cal Z}_2=\Gamma^2F(S^1)_2=\Gamma^2F(y_0,y_1)$ and $\Gamma^3F(y_0,y_1)\subseteq{\cal B}_2$. Now , by the construction of ${\cal B}_2$, ${\cal B}_2\subseteq\Gamma^3F(y_0,y_1)$ and therefore $$\pi_2(F(S^1))\cong{\cal Z}_2/{\cal B}_2=\Gamma^2F(y_0,y_1)/\Gamma^3F(y_0,y_1)=Z(F(y_0,y_1)/{\cal B}_2).$$ The assertion follows. \begin{remark} The relations in above theorem are not minimal, i.e, many of them can be cancelled out. \end{remark} \begin{proposition} Let ${C'}_{n+1}$ be the subgroup of $F(S^1)_{n+1}$ generated by all commutators given by $[\cdots[y_{i_1}^{\epsilon_1},y_{i_2}^{\epsilon_2}],\cdots,],y_{i_t}^{\epsilon_t}]$ with $\{i_1,\cdots,i_t\}=\{0,1,\cdots,n\}$ and $\epsilon_j=\pm1$. Then $$ NF(S^1)_{n+1}/\Gamma^s\cap NF(S^1)_{n+1}\cong C'_{n+1}/\Gamma^s\cap C'_{n+1} $$ for each $s$, where $\Gamma^s=\Gamma^sF(S^1)_{n+1}$ is the $s$-term in the lower central series of $F(S^1)_{n+1}$. \end{proposition} \noindent{\em Proof.} Notice that $C'_{n+1}\subseteq NF(S^1)_{n+1}$. The induced homomorphism $f_s:C'_{n+1}/\Gamma^s\cap C'_{n+1}\rightarrow NF(S^1)_{n+1}/\Gamma^s\cap NF(S^1)_{n+1}$ is a monomorphism. We check that $f$ is an epimorphism. It suffices to show that, for each $w\in NF(S^1)$, there exists a sequence of elements $\{x_j\}$ so that $x_j\in\Gamma^j\cap C'_{n+1}$ and $wx_1x_2\cdots x_s\in\Gamma^{s+1}$ for each $s$. In fact, if this statement holds, $wx_1x_2\cdots x_{s-1}\equiv 1$ $mod$ $\Gamma^s\cap NF(S^1)_{n+1}$ for each $s$ and $w=(wx_1\cdots x_{s-1})\cdot(x_1x_2\cdots x_{j-1})^{-1}\in C'_{n+1}$ $mod\,\Gamma^s$. Now we construct $x_j$ by induction, which depends on $w$. Notice that, for $n\geq1$, $NF(S^1)_{n+1}\in\Gamma^{n+1}\subseteq\Gamma^2$. Choose $x_j=1$ for $j\leq n$. Suppose that there are $x_1,\cdots,x_{s-1}$ so that $x_j\in\Gamma^j\cap C'_{n+1}$ and $wx_1\cdots x_{s-1}\in\Gamma^s$. Since $C'_{n+1}\subseteq NF(S^1)_{n+1}$, $wx_1\cdots x_{s-1}\in\Gamma^s\cap NF(S^1)_{n+1}$ and $d_j(wx_1\cdots x_{s-1})=1$ for $j>1$. Let $\pi:\Gamma^s\rightarrow\Gamma^s/\Gamma^{s+1}$ be the quotient homomorphism. $\pi(wx_1\cdots x_{s-1})$ is a linear combination of basic Lie products. We claim that $\pi(wx_1\cdots x_{s-1})$ is a linear combination of basic Lie products in which each $y_j$ appears in the Lie product for $0\leq j\leq n$. If not, there exists $j$ so that $\pi(wx_1\cdots x_{s-1})=b+c$, where $b$ is a nontrivial linear combination of basic Lie products in which $y_j$ does not appear and $c$ is a linear combination of basic Lie products in which $y_j$ appears. Now the face homomorphism $d_{j+1}:F(y_0,\cdots,y_n)\rightarrow F(y_0,\cdots,y_{n-1})$ induces a homomorphism $$ {\bar d}_{j+1}:\Gamma^sF(y_0,\cdots,y_n)/\Gamma^{s+1}F(y_0,\cdots,y_n)\rightarrow\Gamma^sF(y_0,\cdots,y_{n-1})/\Gamma^{s+1}F(y_0,\cdots,y_{n-1}) $$ and $$1={\bar d}_{j+1}\pi(wx_1\cdots x_{s-1})={\bar d}_{j+1}(b)+{\bar d}_{j+1}(c)={\bar d}_{j+1}(b).$$ Notice that $$d_{j+1}|_{F(y_0,\cdots,y_{j-1},y_{j+1},\cdots,y_n)}:F(y_0,\cdots,y_{j-1},y_{j+1},\cdots,y_n)\rightarrow F(y_0,\cdots,y_{n-1})$$ is an isomorphism. Thus $b=1$. This contradicts to that $b$ is a nontrivial linear combination of a basis. Thus $\pi(wx_1\cdots x_{s-1})$ is a linear combination of basic Lie products in which all of the $y_j$ appear. By Theorem 5.12 in [MKS,pp.337], there exists $x_s$ in $C'_{n+1}$ so that $\pi(wx_1\cdots x_s)=1$, or $wx_1\cdots x_s\in\Gamma^{s+1}$. The induction is finished now. \\ \par By Corollary 3.6, we have \begin{theorem} In the free group $F(y_0,\cdots,y_n)$, $NF(S^1)_{n+1}=[[<y_0>,\cdots,<y_n>]]$ and ${\cal B}F(S^1)_{n+1}=[[<y_{-1}>,<y_0>,\cdots,<y_n>]]$. \end{theorem} Thus Theorem 4.5 can be rewritten as follows. \begin{theorem} In the free group $F(y_0,\cdots,y_n)$ for $n\geq1$, the center $$ Z(F(y_0,\cdots,y_n)/[[<y_{-1}>,<y_0>,\cdots,<y_n>]])\cong\pi_{n+2}(S^3) $$ \end{theorem} \par By Lemma 4.1 and Proposition 3.5, $Kerd_0=<x_0>=<y_{-1}>$ and therefore $$ {\cal Z}_{n+1}=[[<y_0>,\cdots,<y_n>]]\cap<y_{-1}>. $$ Thus we have \begin{theorem}[Theorem 1.4] In the free group $F(y_0,\cdots,y_n)$ with $n\geq1$, $$ [[<y_0>,\cdots,<y_n>]]\cap<y_{-1}>]]/[[<y_{-1}>,<y_0>,\cdots,<y_n>]]\cong\pi_{n+2}(S^3). $$ \end{theorem} \section{On the Homotopy Groups of $\Sigma K(\pi,1)$} In this section, we give group theoretical descriptions for $\pi_*(\Sigma K(\pi,1))$ for any group $\pi$. The proof of Theorem 1.5 is given in this section, where Theorem 5.9 is Theorem 1.5. We will use the notations defined in Section 3. First we extend our description for $\pi_*(S^2)$ to the case $\pi_*(\vee_{\alpha\in J}S^2)$. Recall that $(\vee_{\alpha\in J}S^1)_0=*$, $(\vee_{\alpha\in J}S^1)_1=\{\sigma_{\alpha},*|\alpha\in J\}$ and $(\vee_{\alpha\in J}S^1)_{n+1}=\{s_n\cdots{\hat s_i}\cdots s_0\sigma_{\alpha},*|\alpha\in J,0\leq i\leq n\}$. Let $x^{(\alpha)}_i$ denote $s_n\cdots{\hat s_i}\cdots s_0\sigma_{\alpha}$. Then $$ F(\vee_{\alpha\in J}S^1)_{n+1}=F(x^{(\alpha)}_0,x^{(\alpha)}_1,\cdots,x^{(\alpha)}_n|\alpha\in J). $$ Let $y^{(\alpha)}_j=x^{(\alpha)}_j\cdot x^{(\alpha)}_{j+1}$ for $0\leq j\leq n-1$ and $y^{(\alpha)}_n=x^{(\alpha)}_n$. By Lemma 4.1, we have \begin{lemma} $F(\vee_{\alpha\in J}S^1)_{n+1}=F(y^{(\alpha)}_j|0\leq j\leq n,\alpha\in J)$ with $$ d_j(y^{(\alpha)}_k)=\cases{y^{(\alpha)}_{k-1}& $j\leq k$, \cr 1& $j=k+1$, \cr y^{(\alpha)}_k & $j>k+1$, \cr} $$ and $$ s_j(y^{(\alpha)}_k)=\cases{y^{(\alpha)}_{k+1}& $j\leq k$, \cr y^{(\alpha)}_k\cdot y^{(\alpha)}_{k+1}& $j=k+1$, \cr y^{(\alpha)}_k & $j>k+1$, \cr} $$ for $0\leq j\leq n+1$, where $y^{(\alpha)}_{-1}=(y^{(\alpha)}_0\cdots y^{(\alpha)}_{n-1})^{-1}$ in $F(\vee_{\alpha\in J}S^1)$ \end{lemma} Let $C^J_{n+1}$ denote the subgroup of $F(\vee_{\alpha\in J}S^1)_{n+1}$ generated by all of the commutators $[y^{(\alpha_1)\epsilon_1}_{i_1},\cdots, y^{(\alpha_t)\epsilon_t}_{i_t}]$ with $\{i_1,\cdots,i_t\}=\{0,1,\cdots,n\}$ as sets, where $\epsilon_j=\pm1$, $\alpha_j\in J $ and the commutator bracket $[\cdots]$ runs over all of the commutator bracket arrangements of weight $t$ for each $t$. \begin{lemma} $C^J_{n+1}\subseteq NF(\vee_{\alpha\in J}S^1)_{n+1}$. \end{lemma} \noindent{\em Proof:} For each $1\leq j\leq n+1$, there exists some $i_s=j-1$. Thus $d_j(y^{(\alpha_s)\epsilon_s}_{i_s})=1$ for some $i_s$ and therefore $$ d_j([y^{(\alpha_1)\epsilon_1}_{i_1},\cdots, y^{(\alpha_t)\epsilon_t}_{i_t}])=1 $$ for each $j>0$. The assertion follows. \begin{lemma} For each $1\leq j\leq n+1$, $$ Kerd_j\cap F(\vee_{\alpha\in J}S^1)_{n+1}=<y^{(\alpha)}_{j-1}|\alpha\in J>, $$ the normal subgroup generated by $y_{j-1}^{(\alpha)}$ with $\alpha\in J$. \end{lemma} \noindent{\em Proof:} By the definition of $d_j$, there is a commutative diagram $$ \hspace{0.0in} \begin{array}{cccccc} & F(y^{(\alpha)}_j|0\leq j\leq n,\alpha\in J) & \stackrel{p}{\rightarrow} & F(y^{(\alpha)}_0\cdots{\hat y^{(\alpha)}_{j-1}}\cdots y_n^{(\alpha)}|\alpha\in J) \\ & d_j\downarrow & &\cong \downarrow {\bar d_j} \\ & F(y^{(\alpha)}_j|0\leq j\leq n-1,\alpha\in J)&\stackrel{=}{\rightarrow} & F(y^{(\alpha)}_j|0\leq j\leq n-1,\alpha\in J) \\ \end{array} $$ where $p$ is the projection and $$ {\bar d_j}y^{(\alpha)}_k=\cases{y^{(\alpha)}_{k-1}& $j\leq k$,\cr y^{(\alpha)}_k& $j>k+1$.\cr} $$ The assertion follows. \begin{theorem} Let $C^J_{n+1}$ be defined as above. Then $$ NF(\vee_{\alpha\in J}S^1)_{n+1}=C^J_{n+1} $$ \end{theorem} \noindent{\em Proof:} By lemma 5.1, each $d_j$ with $j>0$ is a projection homomorphism. Thus, by Theorem 3.5, $$ NF(\vee_{\alpha\in J}S^1)_{n+1}=\bigcap_{j=1}^{n+1}Kerd_j=F({\cal A}(T_0,T_1,\cdots,T_n)) $$ where $T_j=\{y^{(\alpha)}_0,\cdots,{\hat y^{(\alpha)}_j},\cdots,y^{(\alpha)}_n|\alpha\in J\}$. It suffices to show that $${\cal A}(T_0,T_1,\cdots,T_n)\subseteq C^J_{n+1}.$$ This follows from the next lemma. \begin{lemma} For $0\leq j\leq n$, let $W=[y^{(\alpha_1)\epsilon_1}_{i_1},\cdots,y^{(\alpha_t)\epsilon_t}_{i_t}]\in {\cal A}(T_0,T_1,\cdots,T_j)$. Then $$ \{0,1,\cdots,j\}\subseteq\{i_1,\cdots,i_t\}. $$ \end{lemma} \noindent{\em Proof:} The proof is given by induction on $j$ for $0\leq j\leq n$. Notice that, by construction, each element in ${\cal A}(T_0,T_1,\cdots,T_j)$ is written as a certain commutator. If $j=0$, then $$ {\cal A}(T_0)=\{y^{(\alpha)}_0,[[y^{(\alpha)}_0,y^{(\alpha_1)\epsilon_1}_{i_1}],\cdots,],y^{(\alpha_t)\epsilon_t}_{i_t}]\} $$ where $\alpha,\alpha_j\in J$, $\epsilon_j=\pm1$ and $y^{(\alpha_1)\epsilon_1}_{i_1}\cdots y^{(\alpha_t)\epsilon_t}_{i_t}$ runs over all of the reduced words $\not=1$ in $F(y^{(\alpha)}_j|\alpha\in J,1\leq j\leq n)$. Thus the assertion holds for $j=0$. Suppose that the the assertion holds for $j-1$ with $j\leq n$. Recall that $$ T^{(j)}_j=\{W\in{\cal A}(T_0,T_1,\cdots,T_{j-1}|W=[y^{(\alpha_1)\epsilon_1}_{i_1},y^{(\alpha_2)\epsilon_2}_{i_2},\cdots,y^{(\alpha_t)\epsilon_t}_{i_t}],\,with\,y^{(\alpha_s)}_{i_s}\in T_j\}. $$ Notice that $y^{(\alpha_s)}_{i_s}\in T_j\Longleftrightarrow i_s\not=j$. Thus, for $W=[y^{(\alpha_1)\epsilon_1}_{i_1},y^{(\alpha_2)\epsilon_2}_{i_2},\cdots,y^{(\alpha_t)\epsilon_t}_{i_t}]\in{\cal A}(T_0,\cdots,T_{j-1})-T^{(j)}_j$, $j\in\{i_1,\cdots,i_t\}$. By induction, $\{0,\cdots,j-1\}\subseteq\{i_1,\cdots,i_t\}$. Hence $$\{0,1,\cdots,j\}\subseteq\{i_1,i_2,\cdots,i_t\}$$ for any $W=[y^{(\alpha_1)\epsilon_1}_{i_1},y^{(\alpha_2)\epsilon_2}_{i_2},\cdots,y^{(\alpha_t)\epsilon_t}_{i_t}]\in{\cal A}(T_0,\cdots,T_{j-1}-T^{(j)}_j$. Recall that, by construction, $$ {\cal A}(T_0,\cdots, T_j)={\cal A}_{T^{(j)}_j}. $$ $$\{0,1,\cdots,j\}\subseteq\{i_1,i_2,\cdots,i_t\}$$ for any $W=[y^{(\alpha_1)\epsilon_1}_{i_1},y^{(\alpha_2)\epsilon_2}_{i_2},\cdots,y^{(\alpha_t)\epsilon_t}_{i_t}]\in{\cal A}(T_0,\cdots,T_{j})$. This completes the proof. \begin{theorem} $\pi_{n+2}(\vee_{\alpha\in J}S^2)$ is isomorphic to the center of the group with generators $$ y^{(\alpha)}_0,y^{(\alpha)}_1,\cdots,y^{(\alpha)}_n $$ for $\alpha\in J$. and relations $$ [y^{(\alpha_1)\epsilon_1}_{i_1}, y^{(\alpha_2)\epsilon_2}_{i_2},\cdots,y^{(\alpha_t)\epsilon_t}_{i_t}] $$ with $\{i_1,i_2,\cdots,i_t\}=\{-1,0,1,\cdots,n\}$ as sets, where the indices $i_j$ can be repeated, $\epsilon_j=\pm1$, $\alpha_j\in J$ and the commutator bracket runs over all of the commutator bracket arrangements of weight $t$ for each $t$. \end{theorem} \noindent{\em Proof:} Notice that $\pi_{n+2}(\vee_{\alpha\in J} S^2)\cong\pi_{n+1}(F(\vee_{\alpha\in J}S^1))$. By the above theorem, ${\cal B}_{n+1}$ is generated by $$ [y^{(\alpha_1)\epsilon_1}_{i_1}, y^{(\alpha_2)\epsilon_2}_{i_2},\cdots,y^{(\alpha_t)\epsilon_t}_{i_t}] $$ By Theorem 2.8, the assertion holds for $n\geq1$. For $n=0$, ${\cal B}_1=\Gamma^2(F(y^{(\alpha)}_0|\alpha\in J))$ and $$ F(y^{(\alpha)}_0|\alpha\in J)/{\cal B}_1\cong \oplus_{\alpha\in J}{\bf Z}\pi_2(\vee_{\alpha\in J}S^2). $$ The assertion holds for all of $n$.\\ \par For the general case, we need a simplicial group construction. \begin{definition} Let $G$ be a simplicial group and let $X$ be a pointed simplicial set with a point $*$. The simplicial group {\bf $F^G(X)$} is defined by setting $$ F^G(X)_n=\coprod_{x\in X_n}(G_n)_x, $$ the free product, modulo the relations $(G_n)_*$, where $(G_n)_x$ is a copy of $G_n$. The faces and degeneracies homomorphisms in $F^G(X)$ is given in the canonical way by the universal property of the coproduct in the category of groups and group homomorphisms. \end{definition} \begin{lemma}[Ca, Theorem 9, pp.88] Let $G$ be a simplicial group and let $X$ be a pointed simplicial set. Then the goemetric realization $|F^G(X)|$ is homotopy equivalent to $\Omega(B|G|\wedge|X|)$. \end{lemma} A generalization of this lemma by using fibrewise simplicial groups is given in [Wu1]. \begin{theorem}[Theorem 1.5] Let $\pi$ be any group and let $\{x^{(\alpha)}|\alpha\in J\}$ be a set of generators for $\pi$. Then, for $n\not=1$, $\pi_{n+2}(\Sigma K(\pi,1))$ is isomorphic to the center of the quotient group of the free product $${\coprod^{groups}}_{0\leq j\leq n}(\pi)_j$$ modulo the relations $$ [y^{(\alpha_1)\epsilon_1}_{i_1},y^{(\alpha_2)\epsilon_2}_{i_2},\cdots,y^{(\alpha_t)\epsilon_t}_{i_t}] $$ with $\{i_1,i_2,\cdots,i_t\}=\{-1,0,1,\cdots,n\}$ as sets, where $(\pi)_j$ is a copy of $\pi$ with generators $\{x^{(\alpha)}_j|\alpha\in J\}$, $\epsilon_j=\pm1$, $y^{(\alpha)}_{-1}={x^{(\alpha)}_0}^{-1}$, $y_j=x^{(\alpha)}_j{x^{(\alpha)}_{j+1}}^{-1}$ for $1\leq j\leq n-1$ and $y^{(\alpha)}_n=x^{(\alpha)}_n$, the commutator bracket $[\cdots]$ runs over all of the commutator bracket arrangment of weight $t$ for each $t$. \end{theorem} \noindent{\em Proof:} Since $\{x^{(\alpha)}|\alpha\in J\}$ is a set of generators for $\pi$, $F(x^{(\alpha)}|\alpha\in J)\rightarrow\pi $ is an epimorphism. Thus $$ F(\vee_{\alpha\in J}S^1)\cong F^{F(x^{(\alpha)}|\alpha\in J)}(S^1)\rightarrow F^{\pi}(S^1) $$ is an epimorphism. Hence $$ NF(\vee_{\alpha\in J}S^1)\rightarrow NF^{\pi}(S^1) $$ and $$ {\cal B}F(\vee_{\alpha\in J}S^1)\rightarrow {\cal B}F^{\pi}(S^1) $$ are epimorphisms. The assertion follows from Theorem 5.6, Lemma 5.7 and Proposition 2.8. \section{Applications} In this section, we consider Cohen's $K$-construction. Our descriptions for the homotopy groups of $3$-sphere gives a calculation for the $K$-construction of $1$-sphere. \begin{definition} Let $X$ be set. The group ${\bf K(X)}$ is defined to the the quotient group of the free group $F(X)$ modulo the normal subgroup generated by all of the commutators $[[x_1,x_2],\cdots,],x_t]$ with $x_i\in X$ and $x_i=x_j$ for some $1\leq i<j\leq t$ Now let $S$ be a pointed simplicial set. The simplicial group ${\bf K(S)}$ is defined to be the quotient simplicial group of $F(S)$ modulo the normal simplicial subgroup generated by all of the commutators $[[x_1,x_2],\cdots,],x_t]$ with $x_i\in S$ and $x_i=x_j$ for some $1\leq i<j\leq t$, where $F(S)$ is the Milnor's $F(K)$-construction for the simplicial set $S$. \end{definition} \begin{definition} The group $Lie(n)$ is the elments of weight $n$ in the Lie algebra $Lie(x_1,x_2,\cdots,x_n)$ which is the quotient Lie algebra of the free Lie algebra $$L(x_1,x_2,\cdots,x_n)$$ over ${\bf Z}$ modulo the two sided Lie ideal generated by the Lie elements $[[x_{i_1},x_{i_2}],\cdots,],x_{i_t}]$ with $i_l=i_k$ for some $1\leq l<k\leq t$. \end{definition} The following lemmas are due to Fred Cohen. \begin{lemma}[Co] $\Gamma^qK(x_1,x_2,\cdots,x_n)=\{1\}$ for $q\geq n$ and $$\Gamma^nK(x_1,x_2,\cdots,x_n)\cong Lie(n),$$ where $\Gamma^qG$ is the $q$-th term in the lower central series for the group $G$ starting with $\Gamma^1G=G$. \end{lemma} \begin{lemma}[Co] In the group $K(x_1,x_2,\cdots, x_n)$, the normal subgroup grnerated by $x_j$ is abelian for each $1\leq j\leq n$. \end{lemma} \begin{lemma}[Co] The set $\{[x_1,x_{\sigma(2)},\cdots,x_{\sigma(n)}]|\sigma\in\Sigma_{n-1}\}$ is a ${\bf Z}$-basis for $Lie(n)$, where $\Sigma_{n-1}$ acts on $\{2, 3,\cdots,n\}$ by permutation. \end{lemma} Recall that the simplicial $1$-sphere $S^1$ is a free simplicial set generated by a $1$-simplex $\sigma$. More precisely, $S^1_0=\{*\}$, $S^1_1=\{\sigma,*\}$ and $S^1_{n+1}=\{*,s_n\cdots{\hat s_j} \cdots s_0\sigma|0\leq j\leq n\}$. Let $x_i$ denote $s_n\cdots {\hat s_j}\cdots s_0\sigma$. Then \begin{lemma} The face functions $d_i:S^1_{n+1}\rightarrow S^1_n$ and the degenarate functions $s_i:S^1_{n+1}\rightarrow S^1_{n+2}$ are as follows: $$ d_ix_j=\cases{x_j&$j<i$\cr x_{j-1}&$j\geq i$\cr} $$ and $$ s_ix_j=\cases{x_j&$j<i$\cr x_{j+1}&$j\geq i$\cr}, $$ where we put $x_{-1}=*$ and $x_n=*$ in $S^1_n$. \end{lemma} \begin{theorem} $\pi_n(K(S^1))$ is isomorphic to $Lie(n)$ \end{theorem} \noindent{\em Proof:} Let $\pi:F(S^1)\rightarrow K(S^1)$ be the quotient homomorphism. Then $NF(S^1)\rightarrow NK(S^1)$ is an epimorphism. Recall that $NF(S^1)_{n+1}$ is generated by all of the commutators $$ [y_{i_1},y_{i_2},\cdots,y_{i_t}] $$ so that $$\{i_1,i_2,\cdots,i_t\}=\{0,1,\cdots,n\}$$ by Theorem 4.3. Thus $ NF(S^1)_{n+1}\subseteq \Gamma^{n+1}F(S^1)_{n+1} $ and therefore $$ NK(S^1)_{n+1}\subseteq\Gamma^{n+1}K(S^1)_{n+1}. $$ Notice that $K(S^1)_{n+1}\cong K(x_0,x_1,\cdots,x_n)$. Thus $\Gamma^{n+1}K(S^1)_{n+1}\cong Lie(n+1)$ and $\Gamma^{n+1}K(S^1)_n=\{1\}$. Thus $$ d_j|_{\Gamma^{n+1}K(S^1)_{n+1}}:\Gamma^{n+1}K(S^1)_{n+1}\rightarrow K(S^1)_n $$ is trivial for each $j\geq0$. And therefore $$ NKS^1)_{n+1}=\Gamma^{n+1}K(S^1)_{n+1}\cong Lie(n+1) $$ with $ d_0:NK(S^1)_{n+1}\rightarrow NK(S^1)_n $ is trivial. the assertion follows. \begin{corollary} Let $\pi:F(S^1)\rightarrow K(S^1)$ be the quotient simplicial homomorphism. Then $$ \pi_*:\pi_n(F(S^1))\rightarrow \pi_n(K(S^1)) $$ is an isomorphism for $n=1,2$ and zero for $n>2$. \end{corollary} Now we consider the Samelson product in $\pi_*(K(S^1))$. Let $x_j$ denote $s_n\cdots{\hat s_j}\cdots s_0\sigma$ in $S^1_{n+1}$. The following lemma follows directly from Lemma 6.6. \begin{lemma} Let $I=(i_1,i_2,\cdots,i_m)$ be a sequence with $i_1<i_2<\cdots<i_m$. Then $ s_I:S^1_{n+1}-*\rightarrow S^1_{n+m+1}-*$ is the composite $$ \{x_0,x_1,\cdots,x_n\}\stackrel{{\bar s_I}}{\rightarrow}\{x_0,x_1,\cdots,{\hat x_{i_1}},\cdots,{\hat x_{i_2}},\cdots,{\hat x_{i_m}},\cdots,x_{n+m}\} $$ $$ \hookrightarrow\{x_0,x_1,\cdots,x_{n+m}\}. $$ where $s_I=s_{i_m}\cdots s_{i_1}$ and ${\bar s_I}$ is the order preserving isomorphism. \end{lemma} Recall that, for $\sigma\in\pi_n(G)$ and $\tau\in\pi_m(G)$, the Samelson product [C1] is defined to be $$ <\sigma,\tau>=\prod_{(a,b)}[s_b\sigma, s_a\tau]^{sign(a,b)}, $$ where $G$ is a simplicial group, $(a,b)=(a_1,\cdots,a_n,b_1,\cdots,b_m)$ runs over all shuffles of $(0,1,\cdots,m+n-1)$,i.e. all permutations, so that $a_1<a_2<\cdots<a_n$, $b_1<b_2<\cdots<b_m$, $sign(a,b)$ is the sign of the permutation $(a,b)$, the order of the product $\prod$ is right lexicographic on $a$ and $s_a=s_{a_n}\cdots s_{a_1}$. \begin{proposition} The Samelson product in $\pi_*(K(S^1))$ is as follows: $$ <[x_{\sigma(0)},x_{\sigma(1)},\cdots,x_{\sigma(n)}],[x_{\tau(0)},x_{\tau(1)},\cdots,x_{\tau(m)}]> $$ $$ =\sum_{(I,J)}sign(I,J)[[x_{i_{\sigma(0)}},\cdots,x_{i_{\sigma(n)}}],[x_{j_{\tau(0)}},\cdots,x_{j_{\tau(m)}}]] $$ for the commutators $$ [x_{\sigma(0)},x_{\sigma(1)},\cdots,x_{\sigma(n)}]\in\pi_{n+1}(KS^1))\cong Lie(n+1) $$ and $$ [x_{\tau(0)},x_{\tau(1)},\cdots,x_{\tau(m)}]\in\pi_{m+1}(K(S^1))\cong Lie(m+1) $$ where $$(I,J)=(i_0,i_1,\cdots,i_n,j_0,j_1,\cdots,j_m)$$ runs over all shuffles of $(0,1,\cdots,m+n+1)$ so that $i_0<i_1<\cdots<i_n$, $j_0<j_1<\cdots<j_m$, $sign(I,J)$ is the sign of the permutation $(I,J)$, $\sigma\in\Sigma_{n+1}$ acts on $\{0,1,\cdots,n\}$ and $\tau\in\Sigma_{m+1}$ acts on $\{0,1,\cdots,m\}$ \end{proposition} \noindent{\em Proof:} Notice that $ \{x_0,\cdots,{\hat x_{j_0}},\cdots,{\hat x_{j_m}},\cdots,x_{n+m+1}\}=\{x_{i_0},\cdots,x_{i_n}\} $ and $ {\bar s_J}:\{x_0,\cdots,x_n\}\rightarrow\{x_{i_0},\cdots,x_{i_n}\} $ is an ordered isomorphism. $$ s_J([x_{\sigma(0)},x_{\sigma(1)},\cdots,x_{\sigma(n)}])=[x_{i_{\sigma(0)}},\cdots,x_{i_{\sigma(n)}}] $$ and $$ s_I([x_{\tau(0)},x_{\tau(1)},\cdots,x_{\tau(m)}])=[x_{j_{\tau(0)}},\cdots,x_{j_{\tau(m)}}]. $$ The assertion follows. \begin{definition} A simplicial group is {\bf minimal} if it is also a minimal simplicial set. \end{definition} Recall that a simplicial group $G$ is minimal if and only if the Moore chain complex $NG$ is minimal [C2]. \begin{theorem} The simplicial group $K(S^1)$ is the universal minimal simplicial quotient simplicial group of $F(S^1)$ in the following sense:\\ (1). $K(S^1)$ is a minimal simplicial group.\\ (2). Let $G$ be a minimal simplicial group. Then every simplicial homomorphism $f:F(S^1)\rightarrow G$ factors through $K(S^1)$. \end{theorem} \noindent{\em Proof:} By inspecting the proof of Theorem 6.7, $K(S^1)$ is a minimal simplicial group. The assertion (2) follows from the following statement.\\ \par \noindent{\em Statement:} $K(S^1)$ is the quotient simplicial group of $F(S^1)$ modulo the normal subsimplicial group generated by the boundaries.\\ \par Let $H$ denote the kernel of the quotient map $p:F(S^1)\rightarrow K(S^1)$ and let $\overline{B}$ denote the normal subsimplicial group of $F(S^1)$ generated by the boundaries ${\cal B}F(S^1)$. Notice that $K(S^1)$ is a minimal simplicial group. Thus $\overline{B}$ is contained in $H$. Let $Q$ denote the quotient simplicial group $F(S^1)/\overline{B}$. Then $Q$ is a minimal simplicial group. By Proposition 2.6, there is a central extension $$ 0\rightarrow K(\pi_{n+1}Q, n+1)\rightarrow P_{n+1}Q\rightarrow P_nQ\rightarrow0, $$ where $P_nQ$ is the n-th Moore-Postnikov system of $Q$. Notice that $P_1Q=K(\pi_Q,1)=K({\bf Z},1)$. Thus $\Gamma^{n+2}P_{n+1}Q=1$ by induction on $n$. Notice that $Q_{n+1}\cong (P_{n+1}Q)_{n+1}$. Thus $\Gamma^{n+2}Q_{n+1}=1$. Now we show that $H$ is contained in $\overline{B}$ by induction on the dimension starting with $H_1=\overline{B}_1=1$. Suppose that $H_n\subseteq\overline{B}_n$ with $n>0$. Notice that $F(S^1)_{n+1}=F(x_0,\cdots,x_n)$ and $K(S^1)_{n+1}=K(x_0,\cdots,x_n)$. Thus $H_{n+1}$ is a normal subgroup of $F(x_0,\cdots,x_n)$ generated by the commutators $$ [[x_{i_1},x_{i_2}],\cdots,x_{i_t}],x_{i_t}] $$ such that $i_p\not=i_q$ for $p<q$. Now consider $W=[[x_{i_1},x_{i_2}],\cdots,x_{i_t}],x_{i_t}]$. If $t\geq n+1$, then $W\in\Gamma^{n+2}F(x_0,\cdots,x_n)$. Thus $W\in\overline{B}_{n+1}$ since $\Gamma^{n+2}Q_{n+1}=1$. \par If $t<n+1$, then there exists an index $j\in\{0,1,\cdots,n\}-\{i_1,\cdots,i_t\}$. Recall that $$ s_ix_k=\cases{x_k&$k<i$,\cr x_{k+1}&$k\geq i$,\cr} $$ for $x_k=s_{n-1}\cdots{\hat s_k}\cdots s_0\sigma\in S^1_n$. Thus $$ s_j[[x_{i'_1}, x_{i'_2}],\cdots,x_{i'_t}],x_{i'_t}]=[[x_{i_1},x_{i_2}],\cdots,x_{i_t}],x_{i_t}], $$ where $i'_k=i_k$ if $i_k<j$ and $i'_k=i_k-1$ if $i_k>j$. By induction, $[[x_{i'_1}, x_{i'_2}],\cdots,x_{i'_t}],x_{i'_t}]\in\overline{B}$. Thus $W=[[x_{i_1},x_{i_2}],\cdots,x_{i_t}],x_{i_t}]\in\overline{B}$. The induction is finished and the assertion follows.\\ \par The simplicial group $K(S^1)$ is homotopy eqivalent to a product of the Eilenberg-Maclane spaces with a different product structure. \begin{proposition} $\Omega K(S^1)$ is an abelian simplicial group. Therefore $K(S^1)$ is homotopy equivalent to a product of the Eilenberg-MacLane spaces as a simplicial set. \end{proposition} \noindent{\em Proof:} Consider $$ d_0:K(x_0,x_1,\cdots,x_n)\rightarrow K(x_0,x_1,\cdots,x_{n-1}) $$ $ d_0(x_0)=1 $ and $ d_0(x_j)=x_{j-1}. $ Thus $Kerd_0\cap K_{n+1}(S^1)\cong <x_0>$ is the normal subgroup generated by $x_0$ which is abelian by Lemma 6.4. The assertion follows.\\ \par In the end of this section, we give some applications of $K(S^1)$ to minimal simplicial groups. \begin{proposition} Let $G$ be a minimal simplicial group such that the abelianlizer $G^{ab}$ is a minimal simplicial group $K(\pi,1)$ for a cyclic group $\pi$. Then $G$ is homotopy equivalent to a product of Eilenberg-Maclane spaces. \end{proposition} \noindent{\em Proof:} Notice that $G_1=\pi$. Let $x$ be a generator for the cyclic group $\pi$ and let $f_x:S^1\rightarrow G$ be a representive map of $x$, i.e, $f_x(\sigma)=x$. Let $g:F(S^1)\rightarrow G$ be the simplicial homomorphism induced by $f_x$. We need a lemma. \begin{lemma} The simplicial homomorphism $g:F(S^1)\rightarrow G$ is simplicial surjection. \end{lemma} \noindent{\em Proof:} It suffices to show that the subsimplicial group, denote by $H$, of $G$ generated by $G_1$ is $G$ itself. This is given by induction on the dimensions starting with $H_1=G_1$. Suppose that $H_{n-1}=G_{n-1}$ with $n>1$. By a result of Condule [see, e.g, Po, Proposition 1, pp.6], $G_n$ is generated by the degenerate images of lower order Moore chain complex terms and $NG_n$. Thus $G_n$ is genereted by $NG_n$ and $H_n$ by induction. Notice that $G$ is a minimal simplicial group. Thus $NG_n={\cal Z}G_n$, the cycles. By Proposition 2.1, ${\cal Z}G_n$ is contained in the center of $G_n$. Thus $H_n$ is a normal subgroup of $G_n$ and the composite $\phi:\pi_nG\cong{\cal Z}G_n\rightarrow G_n\rightarrow G_n/H_n$ is an epimorphism. Thus $G_n/H_n$ is an abelian group and so the quotient homomorphism $G_n\rightarrow G_n/H_n$ factors through $G_n^{ab}$. Notice that $G^{ab}=K(\pi,1)$. Thus $G^{ab}_n\cong G^{ab}_n/{\cal Z}G_n$ and so $\phi:\pi_nG\rightarrow G_n/H_n$ is trivial. Thus $G_n/H_n$ is trivial and the assertion follows.\\ \par \noindent{\em Continuation of Proof of Proposition 6.14:} Notice that $G$ is minimal. The simplicial epimorphism $g:F(S^1)\rightarrow G$ factors through $K(S^1)$ by Proposition 6.12. By Proposition 6.13, $\Omega K(S^1)$ is an abelian simplicial group. Thus $\Omega G$ is also an abelian simplicial group. Thus $G$ is homotopy equivalent to a product of Eilenberg-Maclane spaces, which is the assertion.\\ \par The following counter-example for minimal simplicial groups is due to J. W. Milnor (unpublished). \begin{proposition} $\Omega (S^{n+1}[n+1,n+2,n+3])$ does not have a homotopy type of a minimal simplicial group for $n>0$, where $S^{n+1}[n+1,n+2,n+3]$ is the 3-stage Postnikov system by taking the first three nontrivial homotopy groups of $S^{n+1}$. \end{proposition} \noindent{\em Proof:} Suppose that $G$ is a minimal simplicial group such that $G\simeq\Omega(S^{n+1}[n+1,n+2,n+3])$. Let $f:F(S^1)\rightarrow \Omega^{n-1}G$ be a simplicial homomorphism such that $f(\sigma)$ is a generator of $(\Omega^{n-1}G)_1\cong G_n\cong {\bf Z}$. Then $f_*:\pi_j(F(S^1))\rightarrow\pi_j(\Omega^{n-1}G)$ is an isomorphism for $j\leq3$. Notice that $\Omega^{n-1}G$ is also a minimal simplicial group. The simplicial homomorphism $f:F(S^1)\rightarrow\Omega^{n-1}G$ factors through $K(S^1)$. Notice that $\pi_3(F(S^1))\cong\pi_3\Omega^{n-1}G)\cong{\bf Z}/2$ and $\pi_3(K(S^1))\cong Lie(3)\cong{\bf Z}\oplus{\bf Z}$. There is a contradiction and the assertion follows.\\ \par More examples and counter-examples for minimal simplicial groups will be given in [Wu2]. It was know that there are still many counter-examples of two-stage Postnikov systems for minimal simplicial groups [Wu2]. \bigskip
\section{Preliminaries} Let $k$ be a perfect field of characteristic $p>0$ and $X$ a smooth $k$-variety of dimension $n$. By $\Omega_X$ we denote the sheaf of $k$-differentials on $X$ and $\Omega^j_X=\wedge^j \Omega_X$. The (absolute) Frobenius morphism $F:X\rightarrow X$ is the morphism on $X$, which is the identity on the level of points and given by $F^\#(f)=f^p: \O_X(U)\rightarrow F_*\O_X(U)$ on the level of functions. If ${\E F}$ is an $\O_X$-module, then $F_* {\E F}={\E F}$ as sheaves of abelian groups, but the $\O_X$-module structure is changed according to the homomorphism $\O_X\rightarrow F_*\O_X$. \subsection{The Cartier operator} The universal derivation $d:\O_X \rightarrow \Omega_X$ gives rise to a family of $k$-homomorphisms $d^j: \Omega^j_X\rightarrow \Omega^{j+1}_X$ making $\Omega^\bullet_X$ into a complex of $k$-modules which is called the de Rham complex of $X$. By applying $F_*$ to the de Rham complex, we obtain a complex $F_*\Omega^\bullet_X$ of $\O_X$-modules. Let $B^i_X\subseteq Z^i_X\subseteq F_*\Omega^i_X$ denote the coboundaries and cocycles in degree $i$. There is the following very nice description of the cohomology of $F_*\Omega^\bullet_X$ due to Cartier. \begin{thm} There is a uniquely determined graded $\O_X$-algebra isomorphism $$ C^{-1}:\Omega_X^\bullet\rightarrow \cal H^\bullet(F_* \Omega^\bullet_X) $$ which in degree $1$ is given locally as $$ C^{-1}(da)= a^{p-1} da $$ \end{thm} \begin{pf} \cite{Katz}, Theorem 7.2. \end{pf} With some abuse of notation, we let $C$ denote the natural homomorphism $Z^i_X\rightarrow\Omega^i_X$, which after reduction modulo $B^i_X$ gives the inverse isomorphism to $C^{-1}$. The isomorphism $\bar{C}:Z^i_X/B^i_X\rightarrow \Omega^i_X$ is called the Cartier operator. \section{Liftings of Frobenius to $W_2(k)$} \label{flift} There is a very interesting connection between the Cartier operator and liftings of the Frobenius morphism to flat schemes of characteristic $p^2$. This beautiful observation was first made by Mazur in \cite{Maz}. We go on to explore this next. \subsection{Witt vectors of length two} The Witt vectors $W_2(k)$ (\cite{MumCu}, Lecture 26) of length $2$ over $k$ can be interpreted as the set $k\times k$, where multiplication and addition for $a=(a_0, a_1)$ and $b=(b_0, b_1)$ in $W_2(k)$ are defined by $$ a\, b=(a_0\, b_0, a_0^p b_1+ b_0^p a_1) $$ and $$ a+b=(a_0+b_0, a_1+b_1+\sum_{j=1}^{p-1} p^{-1}\binom{p}{j} a_0^j\, b_0^{p-j}) $$ In the case $k={\Bbb Z}/p$, one can prove that $W_2(k)\cong {\Bbb Z}/p^2$. The projection on the first coordinate $W_2(k)\rightarrow k$ corresponds to the reduction $W_2(k)\rightarrow W_2(k)/p\cong k$ modulo $p$. The ring homomorphism $F^{(2)}:W_2(k)\rightarrow W_2(k)$ given by $F^{(2)}(a_0, a_1)=(a_0^p, a_1^p)$ reduces to the Frobenius homomorphism $F$ on $k$ modulo $p$. \subsection{Splittings of the de Rham complex} The previous section shows that there is a canonical morphism $\operatorname{Spec} k\rightarrow \operatorname{Spec} W_2(k)$. Assume that there is a flat scheme $X^{(2)}$ over $\operatorname{Spec} W_2(k)$ such that \begin{equation} \label{modp} X\cong X^{(2)}\times_{\operatorname{Spec} W_2(k)}\operatorname{Spec} k \end{equation} We shall say that the Frobenius morphism $F$ lifts to $W_2(k)$ if there exists a morphism $F^{(2)}:X^{(2)}\rightarrow X^{(2)}$ covering the Frobenius homomorphism $F^{(2)}$ on $W_2(k)$, which reduces to $F$ via the isomorphism (\ref{modp}). When we use the statement that Frobenius lifts to $W_2(k)$ we will always implicitly assume the existence of the flat lift $X^{(2)}$. \begin{thm} \label{split} If the Frobenius morphism on $X$ lifts to $W_2(k)$ then there is a split quasi-isomorphism $$ 0 @>>> \bigoplus_{0\leq i}\Omega^i_X[-i] @>\sigma>> F_* \Omega^\bullet_X $$ \end{thm} \begin{pf} For an affine open subset $\operatorname{Spec} A^{(2)}\subseteq X^{(2)}$ there is a ring homomorphism $F^{(2)}: A^{(2)}\rightarrow A^{(2)}$ such that $$ F^{(2)}(b)=b^p + p\cdot \varphi(b) $$ where $\varphi: A^{(2)}\rightarrow A=A^{(2)}/ p A^{(2)}$ is some function and $p\, \cdot: A\rightarrow A^{(2)}$ is the $A^{(2)}$-homomorphism derived from tensoring the short exact sequence of $W_2(k)$-modules $$ \CD 0 @>>> p\,W_2(k)@>>> W_2(k) @>p\,\cdot>> p\,W_2(k) @>>> 0 \endCD $$ with the flat $W_2(k)$ module $A^{(2)}$ identifying $A\cong A^{(2)}/p A^{(2)}$ with $p\, A^{(2)}$. We get the following properties of $\varphi$: \begin{align*} \varphi(a+b)&=\varphi(a)+\varphi(b)-\sum_{j=1}^{p-1} p^{-1} \binom{p}{j} \bar{a}^j \bar{b}^{p-j} \\ \varphi(a\, b)&=\bar{a}^p \varphi(b)+ \bar{b}^p \varphi(a) \end{align*} where $\bar{\cdot}$ means reduction $\operatorname{mod} p$. Now it follows that $$ a\mapsto a^{p-1} da+ d \varphi(\tilde{a}) $$ where $\tilde{a}$ is any lift of $a$, is a well defined derivation $\delta:A\rightarrow Z^1_{\operatorname{Spec} A} \subset F_*\Omega^1_{\operatorname{Spec} A}$, which gives a homomorphism $\varphi:\Omega_{\operatorname{Spec} A}^1\rightarrow Z^1_{\operatorname{Spec} A} \subset F_* \Omega^1_{\operatorname{Spec} A}$. This homomorphism can be extended via the algebra structure to give an $A$-algebra homomorphism $\sigma: \oplus_i \Omega_{\operatorname{Spec} A}^i \rightarrow Z^\bullet_{\operatorname{Spec} A}\subseteq F_*\Omega_{\operatorname{Spec} A}^\bullet$, which composed with the canonical homomorphism $Z^\bullet_{\operatorname{Spec} A}\rightarrow \cal H^\bullet(F_* \Omega^\bullet_{\operatorname{Spec} A})$ gives the inverse Cartier operator. Since an affine open covering $\{\operatorname{Spec} A^{(2)}\}$ of $X^{(2)}$ gives rise to an affine open covering $\{\operatorname{Spec} A^{(2)}/p A^{(2)}\}$ of $X$, we have proved that $\sigma$ is a quasi-isomorphism of complexes inducing the inverse Cartier operator on cohomology. Now we give a splitting homomorphism of $\sigma_i:\Omega^i_X\rightarrow F_*\Omega^i_X$. Notice that $\sigma_0:\O_X\rightarrow F_*\O_X$ is the Frobenius homomorphism and that $\sigma_i$ ($i>0$) splits $C$ in the exact sequence $$ \CD 0 @>>> B^i_X @>>> Z^i_X @>C>> \Omega^i_X @>>> 0 \endCD $$ The natural perfect pairing $\Omega_X^i\otimes \Omega^{n-i}_X\rightarrow \Omega^n_X$ gives an isomorphism between ${\cal Hom}_X(\Omega^{n-i}_X, \Omega_X^n)$ and $\Omega^i_X$. It is easy to check that the homomorphism $$ F_*\Omega^i_X\rightarrow {\cal Hom}_X(\Omega^{n-i}_X, \Omega_X^n) \cong \Omega^i_X $$ given by $\omega\mapsto\varphi(\omega)$, where $\varphi(\omega)(z)= C(\sigma_{n-i}(z)\wedge \omega)$, splits $\sigma_i$. \end{pf} \subsection{Bott vanishing} Let $X$ be a normal variety and let $j$ denote the inclusion of the smooth locus $U\subseteq X$. If the Frobenius morphism lifts to $W_2(k)$ on $X$, then the Frobenius morphism on $U$ also lifts to $W_2(k)$. Define the Zariski sheaf $\tilde{\Omega}^i_X$ of $i$-forms on $X$ as $j_*\Omega^i_U$. Since $\operatorname{codim}(X-U)\geq 2$ it follows (\cite{Loc}, Proposition 5.10) that $\tilde{\Omega}^i_X$ is a coherent sheaf on $X$. \begin{thm} Let $X$ be a projective normal variety such that $F$ lifts to $W_2(k)$. Then $$ \H^s(X, \tilde{\Omega}^r_X\otimes L)=0 $$ for $s>0$ and $L$ an ample line bundle. \end{thm} \begin{pf} Let $U$ be the smooth locus of $X$ and let $j$ denote the inclusion of $U$ into $X$. On $U$ we have by Theorem \ref{split} a split sequence $$ 0\rightarrow \Omega^r_U\rightarrow F_*\Omega^r_U $$ which pushes down to the split sequence ($F$ commutes with $j$) $$ 0\rightarrow \tilde{\Omega}_X^r\rightarrow F_*\tilde{\Omega}_X^r $$ Now tensoring with $L$ and using the projection formula we get injections for $s>0$ $$ \H^s(X, \tilde{\Omega}^r_X\otimes L)\hookrightarrow \H^s(X, \tilde{\Omega}^r_X\otimes L^p) $$ Iterating these injections and using that the Zariski sheaves are coherent one gets the desired vanishing theorem by Serre's theorem. \end{pf} \subsection{Degeneration of the Hodge to de Rham spectral sequence} Let $X$ be a projective normal variety with smooth locus $U$. Associated with the complex $\tilde{\Omega}^\bullet_X$ there is a spectral sequence $$ E_1^{pq}=\H^q(X, \tilde{\Omega}^p_X)\implies \H^{p+q}(X, \tilde{\Omega}^\bullet_X) $$ where $\H^\bullet(X, \tilde{\Omega}^\bullet_X)$ denotes the hypercohomology of the complex $\tilde{\Omega}^\bullet_X$. This is the Hodge to de Rham spectral sequence for Zariski sheaves. \begin{thm} If the Frobenius morphism on $X$ lifts to $W_2(k)$, then the spectral sequence degenerates at the $E_1$-term. \end{thm} \begin{pf} As complexes of sheaves of abelian groups $\tilde{\Omega}^\bullet$ and $F_*\tilde{\Omega}^\bullet$ are the same so their hypercohomology agree. Applying hypercohomology to the split injection (Theorem \ref{split}) $$ \sigma:\bigoplus_{0\leq i}\tilde{\Omega}^i_{X/k}[-i]\rightarrow F_* \tilde{\Omega}^\bullet_X $$ we get \begin{eqnarray*} \sum_{p+q=n} \dim_k E_\infty^{pq}= \dim_k \H^n(X, \tilde{\Omega}^\bullet_X)&=& \dim_k \H^n(X, F_*\tilde{\Omega}^\bullet_X)\geq\\ \sum_{p+q=n} \dim_k \H^q(X, \tilde{\Omega}^p_X)&=& \sum_{p+q=n} \dim_k E_1^{pq} \end{eqnarray*} Since $E_\infty^{pq}$ is a subquotient of $E_1^{pq}$, it follows that $E_\infty^{pq}\cong E_1^{pq}$ so that the spectral sequence degenerates at $E_1$. \end{pf} \section{Toric varieties} In this section we briefly sketch the definition of toric varieties following Fulton \cite{Fulton} and demonstrate how the results of Section~\ref{flift} may be applied. \subsection{Convex geometry} Let $N$ be a lattice, $M = \operatorname{Hom}_{{\Bbb Z}}(N, {\Bbb Z})$ the dual lattice, and let $V$ be the real vector space $V = N \otimes_{{\Bbb Z}} {\Bbb R}$. It is natural to identify the dual space of $V$ with $M \otimes_{{\Bbb Z}} {\Bbb R}$, and we think of $N \subset V$ and $M \subset V^*$ as the subsets of integer points. By a cone in $N$ we will mean a subset $\sigma \subset V$ taking the form $\sigma = \{r_1 v_1 + \dots + r_s v_s ~ | ~ r_i \geq 0 \}$ for some $v_i \in N$. The vectors $v_1, \dots, v_s$ are called generators of $\sigma$. We define the dual cone to be $\sigma^{\vee} = \{ u \in V^* | \forall v \in \sigma: \left< u,v \right> \geq 0 \}$. One may show that $\sigma^{\vee}$ is a cone in $M$. A face of $\sigma$ is any set $\sigma \cap u^{\perp}$ for some $u \in \sigma^{\vee}$. Any face of $\sigma$ is clearly a cone in $N$, generated by the $v_i$ for which $\left< u, v_i \right> = 0$. Now let $\sigma$ be a strongly convex cone in $N$, this means that $\{0\}$ is a face of $\sigma$ or equivalently that no nontrivial subspace of $V$ is contained in $\sigma$. We define $S_{\sigma}$ to be the semi group $\sigma^{\vee} \cap M$. Since $\sigma^{\vee}$ is a cone in $M$, $S_{\sigma}$ is finitely generated. \subsection{Affine toric varieties} If $k$ is any commutative ring the semigroup ring $k[S_{\sigma}]$ is a finitely generated commutative $k$-algebra, and $U_{\sigma} = \operatorname{Spec} k[S_{\sigma}]$ is an affine scheme of finite type over $k$. Schemes of this form are called affine toric schemes. \subsection{Glueing affine toric varieties} Let $\tau = \sigma \cap u^{\perp}$ be a face of $\sigma$. One may assume that $u \in S_{\sigma}$. Then it follows that $S_{\tau} = S_{\sigma} + {\Bbb Z}_{\geq 0} \cdot (-u)$, so that $k[S_{\tau}] = k[S_{\sigma}]_{u}$. In this way $U_{\tau}$ becomes a principal open subscheme of $U_{\sigma}$. This may be used to glue affine toric schemes together. We define a fan in $N$ to be a nonempty set $\Delta$ of strongly convex cones in $N$ satisfying that the faces of any cone in $\Delta$ are also in $\Delta$ and the intersection of two cones in $\Delta$ is a face of each. The affine varieties arising from cones in $\Delta$ may be glued together to form a scheme $X_k(\Delta)$ as follows. If $\sigma, \tau \in \Delta$, then $\sigma \cap \tau \in \Delta$ is a face of both $\tau$ and $\sigma$, so $U_{\sigma \cap \tau}$ is isomorphic to open subsets $U_{\sigma\tau}$ in $U_{\sigma}$ and $U_{\tau\sigma}$ in $U_{\tau}$. Take the transition morphism $\phi_{\sigma\tau} : U_{\sigma\tau} \rightarrow U_{\tau\sigma}$ to be the one going through $U_{\sigma \cap \tau}$. A scheme $X_k(\Delta)$ arising from a fan $\Delta$ in some lattice is called a toric scheme. \subsection{Liftings of the Frobenius morphism on toric varieties} Let $X = X_k(\Delta)$ be a toric scheme over the commutative ring $k$ of characteristic $p > 0$. We are going to construct explicitly a lifting of the absolute Frobenius morphism on $X$ to $W = W_2(k)$. Define $X^{(2)}$ to be $X_W(\Delta)$. Since all the rings $W[S_{\sigma}]$ are free $W$-modules, this is clearly a flat scheme over $W_2(k)$. Moreover, the identities $W[S_{\sigma}] \otimes_W k \cong k[S_{\sigma}]$ immediately give an isomorphism $X^{(2)} \times_{\operatorname{Spec} W} \operatorname{Spec} k \cong X$. For $\sigma \in \Delta$, let $F_{\sigma}^{(2)} : W[S_{\sigma}] \rightarrow W[S_{\sigma}]$ be the ring homomorphism extending $F^{(2)} : W \rightarrow W$ and mapping $u \in S_{\sigma}$ to $u^p$. It is easy to see that these maps are compatible with the transition morphisms, so we may take $F^{(2)} : X^{(2)} \rightarrow X^{(2)}$ to be the morphism which is defined by $F_{\sigma}^{(2)}$ locally on $\operatorname{Spec} W[S_{\sigma}]$. This gives the lift of $F$ to $W_2(k)$ and completes the construction. \subsection{Bott vanishing and the Danilov spectral sequence} Since toric varieties are normal we get the following corollary of Section \ref{flift}: \begin{thm} Let $X$ be a projective toric variety over $k$. Then $$ \H^q(X, \tilde{\Omega}^p_X\otimes L)=0 $$ where $q>0$ and $L$ is an ample line bundle. Furthermore the Danilov spectral sequence $$ E_1^{pq}=\H^q(X, \tilde{\Omega}^p_X)\implies \H^{p+q}(X, \tilde{\Omega}^\bullet_X) $$ degenerates at the $E_1$-term. \end{thm} \begin{remark} One may use the above to prove similar results in characteristic zero. The key issue is that we have proved that Bott vanishing and degeneration of the Danilov spectral sequence holds in any prime characteristic. \end{remark} \section{Flag varieties} In this section we generalize Paranjape and Srinivas result on non-lifting of Frobenius on flag varieties not isomorphic to $\P^n$. The key issue is that one can reduce to flag varieties with rank $1$ Picard group. In many of these cases one can exhibit ample line bundles with Bott non-vanishing. We now set up notation. Let $G$ be a semisimple algebraic group over $k$ and fix a Borel subgroup $B$ in $G$. Recall that (reduced) parabolic subgroups $P\supseteq B$ are given by subsets of the simple root subgroups of $B$. These correspond bijectively to subsets of nodes in the Dynkin diagram associated with $G$. A parabolic subgroup $Q$ is contained in $P$ if and only if the simple root subgroups in $Q$ is a subset of the simple root subgroups in $P$. A maximal parabolic subgroup is the maximal parabolic subgroup not containing a specific simple root subgroup. We shall need the following result from the appendix to \cite{MeSri} \begin{prop} \label{splitimplieslift} If the sequence $$ 0 @>>> B^1_X @>>> Z^1_X @>C>> \Omega^1_X @>>> 0 $$ splits, then the Frobenius morphism on $X$ lifts to $W_2(k)$. \end{prop} We also need the following fact derived from (\cite{Hartshorne}, Proposition II.8.12 and Exercise II.5.16(d)) \begin{prop} \label{diffilt} Let $f:X\rightarrow Y$ be a smooth morphism between smooth varieties $X$ and $Y$. Then for every $n\in {\Bbb N}$ there is a filtration $F^0\supseteq F^1 \supseteq \dots$ of $\Omega^n_X$ such that $$ F^i/F^{i+1}\cong f^*\Omega_Y^i\otimes\Omega_{X/Y}^{n-i} $$ \end{prop} \begin{lemma} \label{fibrlemma} Let $f:X\rightarrow Y$ be a surjective, smooth and projective morphism between smooth varieties $X$ and $Y$ such that the fibers have no non-zero global $n$-forms, where $n>0$. Then there is a canonical isomorphism $$ \Omega_Y^\bullet\rightarrow f_*\Omega_X^\bullet $$ and a splitting $\sigma:\Omega^1_X\rightarrow Z^1_X$ of the Cartier operator $C: Z^1_X\rightarrow \Omega^1_X$ induces a splitting $f_*\sigma:\Omega^1_Y\rightarrow Z^1_Y$ of $C:Z^1_Y\rightarrow \Omega^1_Y$. \end{lemma} \begin{pf} Notice first that $\O_Y\rightarrow f_*\O_X$ is an isomorphism of rings as $f$ is projective and smooth. The assumption on the fibers translates into $f_*\Omega_{X/Y}^n\otimes k(y)\cong \H^0(X_y, \Omega^n_{X_y})=0$ for geometric points $y\in Y$, when $n>0$. So we get $f_*\Omega_{X/Y}^n=0$ for $n>0$. By Proposition \ref{diffilt} this means that all of the natural homomorphisms $\Omega_Y^n\rightarrow f_*\Omega^n_X$ induced by $\O_Y\rightarrow f_*\O_X\rightarrow f_*\Omega^1_X$ are isomorphisms giving an isomorphism of complexes $$ \CD 0 @>>> \O_Y @>>> \Omega^1_Y @>>> \Omega^2_Y @>>> \dots \\ @. @VVV @VVV @VVV \\ 0 @>>> f_*\O_X @>>> f_*\Omega^1_X @>>> f_*\Omega^2_X @>>> \dots \endCD $$ This means that the middle arrow in the commutative diagram $$ \CD 0 @>>> B^1_Y @>>> Z^1_Y @>C>> \Omega_Y @>>> 0\\ @. @VVV @VVV @VVV \\ 0 @>>> f_* B^1_X @>>> f_* Z^1_X @>f_* C>> f_*\Omega_X @>>> 0\\ \endCD $$ is an isomorphism and the result follows. \end{pf} \begin{cor} \label{maxparnonlift} Let $Q\subseteq P$ be two parabolic subgroups of $G$. If the Frobenius morphism on $G/Q$ lifts to $W_2(k)$, then the Frobenius morphism on $G/P$ lifts to $W_2(k)$. \end{cor} \begin{pf} It is well known that $G/Q\rightarrow G/P$ is a smooth projective fibration, where the fibers are isomorphic to $Z=P/Q$. Since $Z$ is a rational projective smooth variety it follows from (\cite{Hartshorne}, Exercise II.8.8) that $\H^0(Z, \Omega^n_Z)=0$ for $n>0$. Now the result follows from Lemma \ref{fibrlemma} and Proposition \ref{splitimplieslift}. \end{pf} In specific cases one can prove using the ``standard'' exact sequences that certain flag varieties do not have Bott vanishing. We go on to do this next. Let $Y$ be a smooth divisor in a smooth variety $X$. Suppose that $Y$ is defined by the sheaf of ideals $I\subseteq \O_X$. Then (\cite{Hartshorne}, Proposition II.8.17(2) and Exercise II.5.16(d)) gives for $n\in {\Bbb N}$ an exact sequence of $\O_Y$-modules $$ 0\rightarrow \Omega^{n-1}_Y\otimes I/I^2\rightarrow \Omega^n_X\otimes\O_Y\rightarrow \Omega^n_Y\rightarrow 0 $$ From this exact sequence and induction on $n$ it follows that $\H^0(\P^n, \Omega^j_{\P^n}\otimes\O(m))=0$, when $m\leq j$ and $j>0$. \subsection{Quadric hypersurfaces in $\P^n$} \label{quadric} Let $Y$ be a smooth quadric hypersurface in $\P^n$, where $n\geq 4$. There is an exact sequence $$ 0\rightarrow \O_Y(1-n)\rightarrow \Omega^1_{\P^n}\otimes \O(3-n)\otimes\O_Y \rightarrow \Omega^1_Y\otimes \O_Y(3-n)\rightarrow 0 $$ From this it is easy to deduce that $$ \H^{n-2}(Y, \Omega^1_Y\otimes \O_Y(3-n))\cong \H^1(Y, \Omega^{n-2}_Y\otimes \O_Y(n-3))\cong k $$ using that $\H^0(\P^n, \Omega^j_{\P^n}\otimes\O(m))=0$, when $m\leq j$ and $j>0$. \subsection{The incidence variety in $\P^n\times \P^n$} \label{inc} Let $X$ be the incidence variety of lines and hyperplanes in $\P^n\times \P^n$, where $n\geq 2$. Recall that $X$ is the zero set of $x_0 y_0+\dots+x_n y_n$, so that there is an exact sequence $$ 0\rightarrow \O(-1)\times \O(-1)\rightarrow \O_{\P^n}\times\O_{\P^n} \rightarrow \O_X\rightarrow 0 $$ Using K\"unneth it is easy to deduce that $$ \H^{2n-2}(X, \Omega^1_X\otimes \O(1-n)\times \O(1-n))\cong \H^1(X, \Omega^{2n-2}\otimes \O(n-1)\times \O(n-1))\cong k $$ \subsection{Bott non-vanishing for flag varieties} In this section we search for specific maximal parabolic subgroups $P$ and ample line bundles $L$ on $Y=G/P$, such that $$ \H^i(Y, \Omega_Y^j\otimes L)\neq 0 $$ where $i>0$. These are instances of Bott non-vanishing. This will be used in Section \ref{nonlift} to prove non-lifting of Frobenius for a large class of flag varieties. Let $\O(1)$ be the ample generator of $\operatorname{Pic} Y$. By flat base change one may produce examples of Bott non-vanishing for $Y$ for fields of arbitrary prime characteristic by restricting to the field of the complex numbers. This has been done in the setting of Hermitian symmetric spaces, where the cohomology groups $\H^p(Y, \Omega^q\otimes\O(n))$ have been thoroughly investigated by Sato \cite{Sato} and Snow \cite{Snow1}\cite{Snow2}. We now show that these examples exist. In each of the following subsections $Y$ will denote $G/P$, where $P$ is the maximal parabolic subgroup not containing the root subgroup corresponding to the marked simple root in the Dynkin diagram. These flag manifolds are the irreducible Hermitian symmetric spaces. \subsubsection{Type $A$} \label{A} \begin{picture}(30,30)(-20,30) \put( 0,0){\circle*{4}} \put(-4.5,-3){$\times$} \put( 1.5,0){\line(1,0){22}} \put( 25,0){\circle*{4}} \put( 30.6,-0.5){.\,.\,.} \put( 50,0){\circle*{4}} \put( 51.5,0){\line(1,0){22}} \put( 75,0){\circle*{4}} \put( 76.5,0){\line(1,0){22}} \put(100,0){\circle*{4}} \put(105.6,-0.5){.\,.\,.} \put(125,0){\circle*{4}} \put(126.5,0){\line(1,0){22}} \put(150,0){\circle*{4}} \put(145,-3){$\times$} \end{picture} \vskip 2.0truecm If $Y$ is a Grassmann variety not isomorphic to projective space ($Y=G/P$, where $P$ corresponds to leaving out a simple root which is not the left or right most one), one may prove (\cite{Snow1}, Theorem 3.3) that $$ \H^1(Y, \Omega_Y^3\otimes\O(2))\neq 0 $$ \subsubsection{Type $B$} \begin{picture}(30,30)(-20,30) \put( 0,0){\circle*{4}} \put( 1.5,0){\line(1,0){22}} \put( 25,0){\circle{3}} \put( 26.5,0){\line(1,0){22}} \put( 50,0){\circle{3}} \put( 55.6,-0.5){.\,.\,.} \put( 75,0){\circle{3}} \put( 76.5,0){\line(1,0){22}} \put(100,0){\circle{3}} \put(100, 1.5){\line(1,0){25}} \put(100,-1.5){\line(1,0){25}} \put(108.2,-3){$>$} \put(125,0){\circle{3}} \end{picture} \vskip 2.0truecm Here $Y$ is a smooth quadric hypersurface in $\P^n$, where $n\geq 4$ and Bott non-vanishing follows from Section \ref{quadric}. \subsubsection{Type $C$} \begin{picture}(30,30)(-20,30) \put( 0,0){\circle{3}} \put( 1.5,0){\line(1,0){22}} \put( 25,0){\circle{3}} \put( 26.5,0){\line(1,0){22}} \put( 50,0){\circle{3}} \put( 55.6,-0.5){.\,.\,.} \put( 75,0){\circle{3}} \put( 76.5,0){\line(1,0){22}} \put(100,0){\circle{3}} \put(100, 1.5){\line(1,0){25}} \put(100,-1.5){\line(1,0){25}} \put(108.2,-3){$<$} \put(125,0){\circle*{4}} \end{picture} \vskip 2.0truecm By (\cite{Snow2}, Theorem 2.2) it follows that $$ \H^1(Y, \Omega_Y^2\otimes\O(1))\neq 0 $$ \subsubsection{Type $D$} \begin{picture}(30,30)(-20,30) \put( 0,0){\circle*{4}} \put( 1.5,0){\line(1,0){22}} \put( 25,0){\circle{3}} \put( 26.5,0){\line(1,0){22}} \put( 50,0){\circle{3}} \put( 55.6,-0.5){.\,.\,.} \put( 75,0){\circle{3}} \put( 76.5,0){\line(1,0){22}} \put(100,0){\circle{3}} \put(101.5, 0.5){\line(2, 1){22}} \put(125,12){\circle*{4}} \put(101.5,-0.5){\line(2,-1){22}} \put(125,-12){\circle*{4}} \end{picture} \vskip 2.0truecm For the maximal parabolic $P$ corresponding to the leftmost marked simple root, Y=$G/P$ is a smooth quadric hypersurface in $\P^n$, where $n\geq 4$ and Bott non-vanishing follows from Section \ref{quadric}. For the maximal parabolic subgroup corresponding to one of the two rightmost marked simple roots we get by (\cite{Snow2}, Theorem 3.2) that $$ \H^2(Y, \Omega^4_Y\otimes\O(2))\neq 0 $$ \subsubsection{Type $E_6$} \begin{picture}(30,30)(-20,20) \put( 0,0){\circle*{4}} \put( 1.5,0){\line(1,0){22}} \put( 25,0){\circle{3}} \put( 26.5,0){\line(1,0){22}} \put( 50,0){\circle{3}} \put( 50,-1.5){\line(0,-1){22}} \put( 50,-25){\circle{3}} \put( 51.5,0){\line(1,0){22}} \put( 75,0){\circle{3}} \put( 76.5,0){\line(1,0){22}} \put(100,0){\circle*{4}} \end{picture} \vskip2.0truecm By (\cite{Snow2}, Table 4.4) it follows that $$ \H^3(Y, \Omega^5\otimes\O(2))\neq 0 $$ \subsubsection{Type $E_7$} \begin{picture}(30,30)(-20,20) \put( 0,0){\circle{3}} \put( 1.5,0){\line(1,0){22}} \put( 25,0){\circle{3}} \put( 26.5,0){\line(1,0){22}} \put( 50,0){\circle{3}} \put( 50,-1.5){\line(0,-1){22}} \put( 50,-25){\circle{3}} \put( 51.5,0){\line(1,0){22}} \put( 75,0){\circle{3}} \put( 76.5,0){\line(1,0){22}} \put(100,0){\circle{3}} \put(101.5,0){\line(1,0){22}} \put(125,0){\circle*{4}} \end{picture} \vskip 2.0truecm By (\cite{Snow2}, Table 4.5) it follows that $$ \H^4(Y, \Omega^6\otimes\O(2))\neq 0 $$ \subsubsection{Type $G_2$} \label{G} \begin{picture}(30,30)(-20,30) \put( 0,0){\circle*{4}} \put( 0, 1.5){\line(1,0){25}} \put( 1.5,0){\line(1,0){22}} \put( 0,-1.5){\line(1,0){25}} \put(8.2,-3){$<$} \put(25,0){\circle{3}} \end{picture} \vskip2.0truecm Here $Y$ is a smooth quadric hypersurface in $\P^6$ and Bott non-vanishing follows from Section \ref{quadric}. \subsection{Non-lifting of Frobenius for flag varieties} \label{nonlift} We now get the following \begin{thm} Let $Q$ be a parabolic subgroup contained in a maximal parabolic subgroup $P$ in the list \ref{A} - \ref{G}. Then the Frobenius morphism on $G/Q$ does not lift to $W_2(k)$. Furthermore if $G$ is of type $A$, then the Frobenius morphism on any flag variety $G/Q\not\cong \P^m$ does not lift to $W_2(k)$. \end{thm} \begin{pf} If $P$ is a maximal parabolic subgroup in the list \ref{A}-\ref{G}, then the Frobenius morphism on $G/P$ does not lift to $W_2(k)$. By Corollary \ref{maxparnonlift} we get that the Frobenius morphism on $G/Q$ does not lift to $W_2(k)$. In type $A$ the only flag variety not admitting a fibration to a Grassmann variety $\not\cong\P^m$ is the incidence variety. Non-lifting of Frobenius in this case follows from Section \ref{inc}. \end{pf} \begin{remark} The above case by case proof can be generalized to include projective homogeneous $G$-spaces with non-reduced stabilizers. It would be nice to prove in general that the only flag variety enjoying the Bott vanishing property is $\P^n$. We know of no other visible obstruction to lifting Frobenius to $W_2(k)$ for flag varieties than the non-vanishing Bott cohomology groups. \end{remark} \newpage \bibliographystyle{amsplain} \ifx\undefined\leavevmode\hbox to3em{\hrulefill}\, \newcommand{\leavevmode\hbox to3em{\hrulefill}\,}{\leavevmode\hbox to3em{\hrulefill}\,} \fi
\section{Introduction} \setcounter{equation}{0} \setcounter{footnote}{0} One of the striking features of string theory is target-space duality \cite{GPR}. This duality relates space-times of very different nature, that correspond (locally) to the same CFT. In particular, considering a dualization with respect to a non-abelian isometry, the corresponding symmetry admitted by the dual model is non-local, and exists no more as an isometry. From that, one may conclude the non-local nature of possible transformations between the two models. As suggested by Giveon, Rabinovici and Veneziano \cite{GRV}, and proved by \'{A}lvarez, \'{A}lvarez-Gaum\'{e} and Lozano \cite{AAL}, a $\sigma$-model admitting an isometry with vanishing isotropy is related to its dual by a canonical transformation.\\ In the following we derive an explicit change of variables that produces the dual action for models without isotropy. Dealing with the jacobian, we resort to factorization rules for composite and inverse operators' `determinants'. The structure of the paper is as follows: In section {\bf 2} we present and prove a general change of variables relating (locally) the case without isotropy to its dual, classically; the corresponding jacobian is produced in section {\bf 3}, relying on results from section {\bf 4} (which is somewhat independent), where general rules for decomposition of determinants of composite tensors are given. The rules are inferred by the requirement that functional changes of variables be consistent. The global aspects of the case where the isometry group is abelian and the space is (possiblly) curved can be found in section {\bf 5}. Still in that section - some global aspects of the non-abelian simply-connected case are realized. Section {\bf 6} is dedicated to reviewing some of the above results and their significance. \section{The Change of Variables} \setcounter{equation}{0} The main model regarded in this paper is the general case without isotropy studied in \cite{GR},\cite{EGRSV} and presented here briefly. Consider a target space with coordinates $g$ that transform as $g\rightarrow ug$ where $u,g\in G$, ($G$ is some Lie group), and further inert coordinates $x^i$. The general action can be written in the form \begin{eqnarray} S[\inv g \d g,\inv g \bd g,x]\; = \; \i \Big( (\inv g \d g)^a E_{ab}(x)(\inv g \bd g)^b &+ \nonumber \\ (\inv g \d g)^a F^R_{aj}(x)\bar\partial x^j + \partial x^i F^L_{ib}(x)(\inv g \bd g)^b + \partial x^i F_{ij}(x) \bar\partial x^j &- \nonumber \\ \Phi (x)\partial\bar\partial\sigma\Big)\ , \eer{original} where $\sigma$ is the conformal factor and \begin{equation} (\inv g\partial g)^a \equiv tr (\tilde{T}^a \inv g\partial g) \ \ \Leftrightarrow \ \ \inv g\partial g = (\inv g\partial g)^a T_a\ ,\ \ etc. \eeq{gdga} The generators $T_a$, $a=1,...,dim(G)$, obey \begin{equation} [T_a,T_b]=f^c_{ab}T_c \; \; \; \sp , \eeq{TTfT} and the `dual generators' $\tilde{T}^a$ are defined by the condition \begin{equation} tr(T_a \tilde{T}^b)= \delta_{a}^b \; \; \; \sp . \eeq{trTT} To construct the dual model (for review see \cite{GPR} and references therein), one gauges (minimally) the isometry group with gauge fields $A, {\bar A}$ (in complex worldsheet coordinates). These fields are then constrained to be flat by the addition of the term $\lambda_{c}F^{c}(A,{\bar A})$ to the lagrangian, upon integrating out of the lagrange multipliers $\lambda_{c}$. Gauge fixing $g=1$ then gives the action \begin{equation} S[A,{\bar A},x] + \i \lambda_{c} (\partial {\bar A}^{c} - \bar\partial A^{c} + Af^{c}{\bar A} ) \; \; \; , \eeq{fixed} where $S[A,{\bar A},x]$ is (\ref{original}) with $\inv g \d g$ and $\inv g \bd g$ replaced by the {\em independent} fields $A$ and ${\bar A}$, respectively. ({\bf Note:} from now on matrix and vector indices will sometimes be supressed). Finally, (gaussian) integrations over the gauge fields yield the form of the {\em dual} model (in the non-anomalous case) \begin{eqnarray} S_{dual}[\lambda ,x] &= & \i \Big( (\partial \lambda _a-\partial x^iF^{L}_{ia}) N^{ab}(\bar\partial\lambda_b+F^{R}_{bj}\bar\partial x^j) \nonumber \\ & + & \partial x^i F_{ij}\bar\partial x^j - (\Phi -ln \det N)\partial \bar\partial \sigma \Big) \ , \eer{dual} where \begin{equation} N^{ab}\: \equiv \: [ (E+ f^{c} \lambda_c)^{-1}]^{ab}\; \; \; . \eeq{N}In the anomalous case, the only correction to (\ref{N}) is an extra non-local term proportional to $tr T_{a}$ (\cite{GR},\cite{ALVY},\cite{EGRSV}).\\ In this note we present another way to derive the dual model. Namely\, we perform a change of variables in the functional integral from $\{ g,x\}$ to $\{ \lambda ,x\} $ \begin{equation}\begin{array}{ccccc} a)&(\inv g \d g )E(x)\ &=& \partial \lambda - \partial x F^{L}(x)-(\inv g \d g ) f^{c} \lambda_c &\Leftrightarrow \\ b)& \inv g \d g &=& (\partial \lambda - \partial x F^{L}(x))(E(x)+f^{c}\lambda _{c})^{-1}& \err{Generaltrans} ~\footnote{If the structure constants are totally anti-symmetric (i.e. for compact semi-simple group) one also has $\partial (\inv g \lambda g) = ((\inv g\partial g)^a E_{ab} +\partial x F^L_{b})\inv g T^b g$, from which $\lambda$ is derived explicitly.} (see section 5 for the change of variables of opposite chirality). \\ The proof for that runs as follows:\\ Substituting \ref{Generaltrans}$a$ in the first term of (\ref{original}), using \begin{equation} F[\inv g \d g,\inv g \bd g] = 0 \; \; \; \sp , \eeq{flat} and finally substituting \ref{Generaltrans}$b$ one gets the identity \begin{eqnarray} (\inv g \d g)^a E_{ab}(x)(\inv g \bd g)^b +(\inv g \d g )^a F^R_{aj}(x)\bar\partial x^j + \partial x^i F^L_{ib}(x)(\inv g \bd g)^b \partial x^i & &= \nonumber \\ (\partial \lambda _a-\partial x^iF^{L}_{ia}) [(E+ f^{c} \lambda_c)^{-1}]^{ab}(\bar\partial\lambda_b+F^{R}_{bj}\bar\partial x^j)\; + \; \bar\partial(\lambda\inv g \partial g)-\partial(\lambda\inv g \bar\partial g). \eer{identity} Equation (\ref{identity}) relates an action in group variables to an action in the algebra variables, up to a total derivative term which is discussed in section 5. This completes the proof in the level of the lagrangian.\\ Next, we turn to study the jacobian for the transformation (\ref{Generaltrans}). \section{The Jacobian} \setcounter{equation}{0} Denoting both sides of (\ref{Generaltrans})$b$ as ${\cal A}$, their variations with respect to $g$ and $\lambda$ are respectively \begin{equation}\begin{array}{lcccc} LHS \; \; \; & & D^{c}(\inv g \delta g)& =& \partial (\inv g \delta g)^{c} + {\cal A} ^{a} f_{ab}^{c} (\inv g \delta g)^{b} \\ RHS \; \; \; & & \left( -\tilde{N} \tilde{D}(\delta \lambda)\right) ^{c} &= & N^{bc} (\partial \delta \lambda_{b} -{\cal A} ^{a}f_{ab}^{d}\delta \lambda _{d})\; \; \; , \err{variation} where $\:\tilde{}\:$ denotes the functional transpose~\footnote{Notice the similarity between the characterization of the symmetry $ D(\inv g \delta g)=0 $ of the original model - and the corresponding symmetry of its dual: $\tilde{D}(\delta \lambda)=0$. See section 6 for a possible significance of such similarity. The (global) {\em gauge invariant measure} for $\lambda$ is given in ($\ref{variation}_{RHS}$). Its dependence on the background fields is due to the non-local nature of the symmetry as we transform $g\rightarrow\lambda$. Another thing worth mentioning at this point is that (\ref{Generaltrans}) also relates {\em two different field equations} - that of the original model, which looks like: $\partial \bar J (g,x) +\bar\partial J(g,x) =0$ and that of the dual - $F(A(\lambda ,x ),{\bar A} (\lambda ,x))=0$ (see \cite{GR} for the exact forms).}. The required jacobian is thus \begin{equation}\begin{array}{ccccc} J &=&\frac{{\cal D} g}{{\cal D} \lambda }&=& | D^{\scriptscriptstyle -1}\tilde{N}\tilde{D}| \; \; \; , \err{Jacobian} where $|x|$ stands for $\det(x)$. This can be proved to be the ratio \begin{eqnarray} | N| \frac{| D\bar D| }{|D ||\bar D|}\Big| _{{\bar g} =g }&=&| N|\exp \{ -\iv{tr}(\inv g \d g \bar\partial\sigma+\inv g \bd g\partial\sigma )\} \eer{NL} calculated in \cite{EGRSV} to the first order in the conformal factor $\sigma$. These determinants correspond to the changes $F(A(g),{\bar A} ({\bar g} ))\rightarrow g,g\rightarrow A,{\bar g}\rightarrow{\bar A}$ (parametrized as in (\ref{A's})). A general practice for treating `determinants' of inverse and transpose operators is one of the offshoots of the next section, where the equivalence of (\ref{Jacobian}) and (\ref{NL}) follows naturally. \\ One might want to derive the general form of the dual model found in \cite{EGRSV}. To that end, notice that \begin{equation}\begin{array}{ccrclc} 0 &=& tr \int\sigma F(\inv g \d g ,\inv g \bd g )&=&tr\int \sigma\partial (\inv g \bd g ) -tr \int\sigma\bar\partial (\inv g \d g )&\Rightarrow \\ & & tr\int \inv g \bd g \partial\sigma &=&tr\int \inv g \d g \bar\partial\sigma & \err{ganomaly} so that by (\ref{Generaltrans}), (\ref{NL}) becomes \begin{eqnarray} J &=& | N| \exp \{ -\iv{2}\bar\partial\sigma N(\partial \lambda - \partial x F^{L})\} \; \; \; ; \eer{noname} then, by substituting the equations of motion for $\lambda$ \cite{GR}, one obtains the same form for the terms linear in $\sigma$ as in \cite{EGRSV}~\footnote{A quick way to obtain the very result found in \cite{EGRSV} is by using the $\bf \sigma$ {\em dependent} transformation $\inv g \d g= (\partial \lambda - \partial x F^{L}+\partial\sigma tr\: T)N$.}. \section{Factorization of Jacobians} \setcounter{equation}{0} {\em The Motivation}\\ Consider some action $S$ that depends on the group variables $g,{\bar g}$ through the (gauge) fields \begin{equation} A =\inv g \d g \; \; \; , \; \; \; \sp {\bar A}= \bar g^{-1} \bar\partial \bar g \; \; \; \sp , \eeq{A's} which in turn appear in $S$ in combinations $F(A,{\bar A} )=D {\bar A} -\bar\partial A$. When changing variable ${\bar g}\rightarrow {\bar A}$ followed by the change ${\bar A}\rightarrow F$, the {\em total} jacobians multiplying the partition function's integrand are \begin{eqnarray} J& = &(| D | | \bar D | )^{-1} \; \; \; . \eer{ratio} On the other hand, changing $F \rightarrow {\bar g}$ directly, one collects the jacobian $|D \bar D | ^{-1}$. Now, for path integration to be consistent, the results of these two courses of changing variables from ${\bar g}$ to $F$ should be no different in the end. But as we saw in (\ref{NL}), their ratio is non-trivial, at least for groups with traceful structure constants; this is {\em the mixed anomaly} \cite{ALVY}, here in the form of a `multiplicative anomaly' \cite{EGRSV} which seems to violate the functional chain rule.\\ {\em Factorization}\\ To resolve this puzzle~\footnote{The author wishes to thank S. Elitzur for suggesting the direction which led to the formulation and also for the proof (\ref{nonmixed}).} the ghost actions (the variation of which with respect to the conformal factor should give the value of the anomaly) defining the functional determinants are invoked~\footnote{the $\zeta$-function procedure is not defined for jacobians relating two spaces of different types. When both procedures are defined, one might want to prove that they are different by local counterterms at most.}. Let us write the ghost actions in interest: we have $g,\; {\bar g}$ and $F$ that are worldsheet scalars where $A,{\bar A}$ are components of a worldsheet vector; the fermionic ghosts of types $s$ (for scalar) and $v$ (vector) are thus introduced. Changing $A \rightarrow g$, the corresponding jacobian may be written as~\footnote{Such definitions are not Lorentz-invariant and are ill-defined in general \cite{ALVY}. This, however, should not interfere with our argument which is compelled by the chain rule.} \begin{equation}\begin{array}{ccccc} {{\cal D} A}/{{\cal D} g}&=&| D| _{vs}&\equiv & \int {\cal D} v {\cal D} s \exp {\int vDs} \err{D} where the integration in the exponent is over the worldsheet, partial derivatives change to worldsheet covariant ones, and indices of all types are supressed. By partial integration we have \begin{equation}\begin{array}{ccccccc} \int vDs&=&\int s\tilde{D}v &\Rightarrow & | D| _{vs}&= & | \tilde{D} | _{sv} \; \; \; . \err{a} Further, bearing in mind the chain rule, one can factorize and re-merge `determinants' of vector operators` products and derive identities such as: \begin{equation}\begin{array}{cccccc} | O_{1}O_{2}| _{ss} &= &| O_{1}| _{sv} | O_{2}| _{vs} &= &| \tilde{O_{1}}| _{vs}| \tilde{O_{2}}| _{sv} &= \\ | \tilde{O_{1}}\tilde{O_{2}}| _{vv} &=& | (O_{2}O_{1}\tilde{)}| _{vv}&=& | O_{2}O_{1}| _{vv}& \; \; \; . \err{ident1} By considering a change of variables and its inverse change, we also have \begin{equation}\begin{array}{ccccc} | O_{1}^{\scriptscriptstyle -1}O_{2}| _{ss} &= & | O_{1}^{\scriptscriptstyle -1}| _{sv} | O_{2}| _{vs} &=& | O_{1}| _{vs}^{-1} | O_{2}| _{vs} \; \; \; . \err{ident2} These rules for chaining jacobians are easily generalized to jacobians of tensors relating two objects, possibly of different ranks (with respect to worldsheet diffeomorphisms). The basic rules for that are \begin{equation}\begin{array}{cccc} a)\; \; \; &| A| _{r_{1}r_{2}}&=&| \tilde{A} | _{r_{2}r_{1}} \\ b)\; \; \; &| A| _{r_{1}r_{2}} | B| _{r_{2}r_{3}}&=&| AB| _{r_{1}r_{3}}\\ c)\; \; \; &| A^{\scriptscriptstyle -1}| _{r_{1}r_{2}}&=&| A| ^{-1} _{r_{2}r_{1}} \err{rules} with obvious notations. These rules should be correct {\em whenever they correspond to legitimate changes of variables.} \\ Applying these rules to the anomaly (\ref{NL}), it may take the following equivalent forms \begin{eqnarray} \frac{|D \bar D | _{ss}}{| {D}| _{vs}| \bar D| _{vs} } = \frac{|{D}|_{sv}|\bar D| _{vs} }{| {D}| _{vs}| \bar D| _{vs}}= \frac{|D | _{sv}}{ | D | _{vs}} =\frac{|D | _{sv}}{ | \tilde{D} | _{sv}} \nonumber \\ = | \tilde{D}^{\scriptscriptstyle -1}| _{vs} | D | _{sv}= | \tilde{D}^{\scriptscriptstyle -1} D | _{vv}= | D\tilde{D}^{\scriptscriptstyle -1}| _{ss}\; \; \; . \eer{forms} This proves the equality of (\ref{Jacobian}) and (\ref{NL}) in particular. As to the example in the beginning of the section, the corresponding ratio is \begin{eqnarray} \frac{|D\bar D |_{ss}}{|D|_{sv}|\bar D |_{vs}}&=&1 \eer{b} according to (\ref{forms}) or (\ref{rules}).\\ With (\ref{forms}) in mind, let us consider the anomaly in two classes of cases. By definition, we have \begin{eqnarray} \int vDs&=&\int v_{b}\partial s^{b} + v_{c}f^c_{ab}A^a s^b . \eer{D} {\em If $G$ is semisimple,} then $f_{abc}$ is totally antisymmetric. Reshuffeling indices and integrating by parts, (\ref{D}) may be written as \begin{eqnarray} \int -s_{b}\partial v^{b} - s_{b}f^b_{ac}A^a v^c=-\int sDv &\Rightarrow&|D|_{vs} =|D|_{sv} \eer{} $\Rightarrow$ no anomaly, by virtue of (\ref{forms}).\\ {\em If $G$ is solvable}, there exists a triangular basis for the algebra s.t. \begin{equation}\begin{array}{rll} \int vDs=\\ \nonumber \int & v_1\partial s^1 +A^\mu f^1_{\mu 1} v_1 s^1 \\ \nonumber \; \; \; +&v_2\partial s^2 +A^\mu f^1_{\mu 2} v_1 s^2 +A^\mu f^2_{\mu 2} v_2 s^2 \\ \nonumber &\; \; \; .\sp\sp\sp\sp\sp . \sp\sp\sp\sp\sp \; \; \; \sp.\\ \nonumber &\; \; \; .\sp\sp\sp\sp\sp . \sp\sp\sp\sp\sp \spp. \\ \nonumber &\; \; \; .\sp\sp\sp\sp\sp . \sp\sp\sp\sp\sp \spp\; \; \; \sp\; \; \; . \\ \nonumber \; \; \; +& v_N\partial s^N +A^\mu f^1_{\mu N} v_1 s^N+\; \; \; .\; \; \; .\; \; \; . \; \; \; +A^\mu f^N_{\mu N} v_N s^N\: . \err{D2} Integrating over $s_1$ and then over $v_1$, produces the functional determinant \begin{eqnarray} |-\partial+A^\mu f^1_{\mu 1}|_{sv}\; \; \; , \eer{2.5} while setting $v_1$ to zero, by which all of the terms in the second column vanish. Repeating this procedure for $(s^2 , v_2),...,(s^N , v_N)$, we finally get the formula \begin{eqnarray} | D|_{vs} &=& \prod_{k=1}^N |-\partial+A^\mu f^k_{\mu k}|_{sv}=\prod_{k=1}^N |\partial\;{\mathbf -}\; A^\mu f^k_{\mu k}|_{sv}\; \; \; . \eer{DE} Switching $s$ and $v$, we get \begin{eqnarray} &| D|_{sv}\; \; \; =\; \; \; \prod_{k=1}^N |-\partial+A^\mu f^k_{\mu k}|_{vs}=\prod_{k=1}^N |\partial-A^\mu f^k_{\mu k}|_{vs}\; \; \; =&\nonumber\\ &\prod_{k=1}^N |\partial\;{\mathbf +} \; A^\mu f^k_{\mu k}|_{sv}\; \; \; ,& \eer{DD} so the anomaly , which is the ratio of (\ref{DE}) and (\ref{DD}) can be written as a product of chiral anomalies. By Adler and Bardeen \cite{ADBAR}, we conclude that if the anomaly vanishes to first order, it cancels altogether. The condition for that is \cite{EGRSV} \begin{eqnarray} \sum_k f^k_{\mu k}&=&0\; \; \; . \eer{D5} The methods above can also be used for the general case, i.e. a semi-direct product of a semisimple group and a solvable one (e.g. the Lorentz group). However, the general classification of such groups is still a mystery and so is the general rule for factorization of the corresponding covariant derivatives. \section{Global Aspects} \setcounter{equation}{0} {\bf Notations and mathematical tools on compact Riemann surfaces}\\ {\bf 1)} The $z$-component of a one-form $\omega =(\omega_z,\omega_{\bar z})$ can be completed to a full {\em closed} singled-valued one-form of which $\omega_z$ is its $z$-componet, as follows\begin{eqnarray} \tilde \omega =(\omega_z,\tilde \omega_{\bar z})= (\omega_z,\bar\partial\int_{z_0} dz\, \omega_z) = dC\; \; \; \sp (z_0 =const)\; \; \; , \eer{complete} where $C$ is a multivalued function on the surface. One can verify the above on the torus and therefore on every handle of the surface.\\ {\bf 2)} For a closed one-form $Y$, let $Y_{0}$ stand for its zero mode(s) where $Y_{e}$ denotes its exact part, that is, $Y_0\equiv\sum_j \alpha_j\oint_{\aleph_j} Y$ and $Y_e \equiv Y-Y_0$, where $\{\alpha_j\}, j=1,...2{\bf g}$ is the (unique) basis to the space of harmonic differentials on a surface of genus $\bf g$, satisfying $\oint_{\aleph_j} \alpha_k= \delta_{jk}$ and $\{\aleph_j\}$ is a basis to the first homology group of the surface.\\ {\bf 3)} For two closed one-forms $\alpha $ and $\beta $ one has \begin{eqnarray} \int \alpha\wedge \beta &=& \sum _{i=1}^{n} \oint_{a_i}\alpha\oint_{b_i}\beta -\oint_{b_i}\alpha\oint_{a_i}\beta \; \; \; , \eer{riemann} where $\{a_{i},b_{i}|i=1,...,{\bf g}\}$ denote the non-trivial cycles on the $\bf g$-genus surface \cite{Farkas}.\\ {\bf 4)} If $\theta (z,\bar z )$ and $\lambda ( z, \bar z )$ are two multivalued functions on a surface one has \begin{eqnarray} \int d^2z\,(\partial \theta )_0 (\bar\partial \lambda )_e =\int d^2z\, \bar\partial ((\partial \theta )_0\lambda_s ) -\int d^2z\, \bar\partial(\partial \theta )_0\lambda_s =0 -0 =0\; \; \; , \eer{nonmixed} where the subscript $s$ stands for the single valued part, that is, \begin{eqnarray} \lambda_s (z, \bar z) = \int_{z_0}^z \,(d\lambda)_e\; \; \; , \eer{} and thus $ (d\lambda)_e = d\lambda_s$. We deduce\ the decoupling of terms, that are a product of zero modes and exact modes - from the action.\\ {\bf 5)} The equations of motion for the zero modes of a one-form ${\mathbf A}^a=(A,{\bar A})^a$ in actions of the form \begin{eqnarray} {\mathbf S}&=&\i (A M{\bar A} + A N +K{\bar A}) \eer{} where the matrices $M,N$ and $K$ don't depend on $A$, {\em may be} written as \begin{eqnarray} (M{\bar A} +N)_0&=&( AM +K)_0\; \; \; =\; \; \; 0\; \; \; . \eer{zmeof} The zero modes are defined by {\bf 2)} after completing $N$, $K$, $A$ and ${\bar A}$ ({\bf 1)}) and using {\bf 4)} for the decoupling of the exact and the zero modes. \\ {\tt An abelian case}\\ {\em The original model}: Let the original action be \begin{eqnarray} S_{original}&=&\i\partial\theta R^{2}\bar\partial\theta\; \; \; , \eer{abeloriginal} with $R=R(x(z,\bar z))$, $\theta = \theta +2\pi l\; \; \; ,l=const$. In order to dualize (\ref{abeloriginal}) we write an equivalent action \begin{eqnarray} S&=& \Ri\partial\theta R^{2}\bar\partial\theta\; \; \; +\; \; \; (\partial\theta)_0(\bar\partial\lambda)_{0}-(\partial\lambda)_{0}(\bar\partial\theta)_0\; \; \; , \eer{start} where now $\theta$'s holonomies are any $2{\bf g}$ real numbers and $\lambda$ (who's single valued part is yet unspecified) satisfies \begin{eqnarray} \oint_{a_i}(d\lambda)_0 =\frac{4\pi m_{i}}{l} \;,\; \; \; \oint_{b_i}(d\lambda)_0 =\frac{4\pi n_{i}}{l}\; \; \; . \eer{periodicities} To recover $\theta$'s original periodicity, we write the total derivative term in (\ref{start}) as $\frac{i}{4\pi} \int (d\theta)_0\wedge (d\lambda) _0$, which by (\ref{riemann}) equals $\frac{i}{l}\sum_{j}(n_{j} \oint _{a_{j}}d\theta -m_{j} \oint _{b_{j}}d\theta)$. {\em Notice that the zero subscripts in} (\ref{periodicities}) {\em are omittable}. Summing over all possible integer $n_{j}$ and $m_{j}$, constrains $\theta$'s periodicity to be $2\pi l$.\\ {\em The dual model}: Imposing the equations of motion of $(\partial \theta)_0$ and $(\bar\partial \theta)_0$ on (\ref{start}) imposes \begin{eqnarray} (\partial\lambda)_0 - (\partial\theta R^2)_0&=&0\; \; \; . \eer{abelglobal} Therefore, by virtue of {\bf(1)} we define \begin{eqnarray} (\partial\lambda)_e&=&(R^2\partial\theta)_e \; \; \; , \eer{eabelcov} and use (5.10) to write \begin{eqnarray} \partial\lambda&=&R^2\partial\theta \eer{abelcov} and substitute it in the first term of (\ref{start}), which after omitting the zero subscripts form its next terms takes the form \begin{eqnarray} &\Ri\partial\lambda\bar\partial\theta +\partial\theta\bar\partial\lambda-\partial\lambda\bar\partial\theta =\Ri\partial\theta\bar\partial\lambda&\stackrel{\partial \theta=\partial\lambda R^{-2}}{=} \eer{} \begin{eqnarray} &S_{dual}\; \; \; =\; \; \; \Ri\partial\lambda R^{-2}\bar\partial\lambda \; \; \; , & \eer{abeldual} where $\lambda=\lambda+\frac{4\pi}{l}$.\\ This can readily be generalized to (\ref{original}) with $G=U(1)^{dimG}$. To omit the ${\cal R}e$ in (\ref{abeldual}), one may change variables by replacing an integrand of the same form, with a real $\lambda$. The realness issue carries over to the non-abelian case and so should its solution.\\ {\tt A non-abelian case}: Let the original action be \begin{equation}\begin{array}{ccc} a)&\; \; \; S = \i(\inv g \d g)E(\inv g \bd g) =\\ \nonumber b)&\; \; \; \i((\inv g \d g)E)E^{-1}(E(\inv g \bd g)) = \\ \nonumber c)&\; \; \; tr\Ri(g(\inv g \d g) Eg^{-1})_{0}(\bar\partial g g^{-1})_{0}+(g(\inv g \d g) Eg^{-1})_{e}(\bar\partial g g^{-1})_{e}\; \; \; , \err{Nstart} with $G$ simply connected and $(\inv g \d g)E \stackrel{def}{=}((\inv g \d g)E)_b T^b\stackrel{def}{=}(\inv g \d g)^{a}E_{ab}T^b\;$, etc. To derive (\ref{Nstart}c) from (\ref{Nstart}a), we have used the trace cyclicity along with the conclusion from {\bf 4)}.\\ {\em The dual model}\,\, Since all mappings from a Riemann surface to a simply connected group are homeotopic to a point (on the group's manifold), the configurations space of $g$ is continuous and connected, therefore we can invoke the equations of motion of $(Eg^{-1}dg)_0$ together with {\bf5)}, to constrain $g$'s configurations s.t. \begin{eqnarray} ( g^{-1} dg)_0 &=& 0\; \; \; ; \eer{zm} {\em the extent of validity of this step is yet to be examined (elsewhere)}. Substituting $g$ with $g^{-1}$ we can characterize this subspace also by \begin{eqnarray} (gdg^{-1})_0 = - (dg g^{-1})_0 & = & 0\; \; \; , \eer{zm1} so that the first term in (\ref{Nstart}c) decouples from the action. Using {\bf 1)} we define the function $\chi$ by \begin{eqnarray} \partial\chi&=& g(\inv g \d g) Eg^{-1}\; \; \; . \eer{ONcov} Notice that $\chi$'s possible multivalued part decouples from the action and may therefore be taken to be zero. We substitute (\ref{ONcov}) in what's left of (\ref{Nstart}) to give \begin{eqnarray} tr\Ri\partial\chi \bar\partial g g^{-1}\; \; \; . \eer{c} Integrating it by parts and then using the identity $\partial(\bar\partial g g^{-1})=g\bar\partial (g^{-1}\partial g)g^{-1}$ yield \begin{eqnarray} -tr\Ri\chi g\bar\partial (g^{-1}\partial g)g^{-1}\; \; \; . \eer{b} Then, using the trace cyclicity and integrating by parts again we obtain \begin{eqnarray} tr\Ri\inv g \d g\bar\partial\lambda\; \; \; , \eer{a} where $\lambda=g^{-1}\chi g$ is single valued. Finally, we observe that (\ref{ONcov}) may equivalently be written as \begin{eqnarray} \inv g \partial (g\lambda\inv g)g = (\inv g \d g)E \Leftrightarrow \inv g \d g\lambda+\partial\lambda-\lambda\inv g \d g = \inv g \d g E \Leftrightarrow\nonumber \\ \partial\lambda=\inv g \d g(E+f^c\lambda_c)\Leftrightarrow\inv g \d g=\partial\lambda(E+f^c\lambda_c)^{-1}\; , \eer{e} from which we obtain \begin{eqnarray} S_{dual}&=&\i\partial\lambda(E+f^c\lambda_c)^{-1}\bar\partial\lambda\; \; \; , \eer{Ndual} where {\em the $\lambda$'s are single valued.} The total derivative term from the two integrations by parts \begin{eqnarray} tr\frac{i}{2}\Ri d(\chi dg\inv g) \eer{d} vanishes because that $\chi dg\inv g$ is single valued. The generalization to (\ref{original}) is again, straightforward. \section{Concluding Remarks} The constructions presented in this paper imply (once again) that in the case without isotropy, the conformal invariance of an action and that of its dual are equivalent {\em to all orders of} $\alpha '$ (assuming a correct computation of the jacobian, especially if that is non-local).\\ The author's hope is that the change of variables presented in this work will shed some light on the global issues in dual models of isometry groups the mappings to which from Riemann surfaces fall into more interesting homology structure than the one presented here i.e. the trivial one that corresponds to simply-connected groups. \\ The (classical) equality in the case without isotropy (\ref{identity}) between the original and the dual action might seem a surprise, since the former admits a symmetry that seemingly is absent in the latter \cite{OQ}. Further, a check of the change of variables (\ref{Generaltrans}) verifies that transforming $g \rightarrow ug$ induces no change in the dual coordinate $\lambda$. Where has the original symmetry gone? It has gone non-local (cf.\cite{GR}) just to avoid detection by the Killing equation~\footnote{With (\ref{Generaltrans}) in mind, one suspects the existence of some non-local generalization of the Killing equation that is capable of detecting such hidden symmetries, thus providing means by which dualization may be reversed.}. Actually, the very condition for the `smoothness' of the change of variables, that is, for the jacobian to be local, is that the {\em symmetry factors} (see footnote after eq. (\ref{variation})) $|D |$ and $| \tilde{D} |$ before the volume elements ${\cal D} g$ and ${\cal D} \lambda$ respectively should cancel out. It is precisely when this correspondence between the symmetries breaks, that the anomaly occurs. \\ As shown in section {\bf 4}, the factorization approach to reorganize composite jacobians may be of help in tracing and isolating the very generators of the anomaly, as well as in simplifying its ghost action. One might want to generalize that to all Lie algebras and prove the chiral nature of the mixed anomaly in general.\\ \vskip .2in \noindent {\bf Acknowledgements} \vskip .2in \noindent I would like to thank my supervisor A. Giveon for his guidance and for essential help in preparing this paper. I acknowledge S. Elitzur for very useful discussions, and thank A. Babichenko, S. Forste, O. Pelc and G. Sengupta for their help. Special thanks to G. Spivak for her help and support. This work is supported in part by BSF - American-Israel-Binational-Science-Foundation and by the BRF - Basic-Research-Foundation.
\section{#1}} \def
\section{\twebf Introduction} \renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0} The Standard Model (SM) (Weinberg, 1967; Salam, 1969; Glashow {\it et al.}, 1970) of strong and electroweak interactions has been tested with precise experiments at the present colliders (for reviews see, e.g., Altarelli, 1989; Marciano, 1991, 1993a,b; Bethke, 1992; Bethke and Pilcher, 1992; R.\ K.\ Ellis, 1992; Langacker, Luo and Mann, 1992; Brodsky, 1993; \linebreak Sirlin, 1993a,b; Kniehl, 1994a; Soper, 1995). However, the decisive confirmation of the SM or its modification is still ahead. One awaits with great interest precise measurements from LEP, SLC, HERA, Fermilab Tevatron, etc. The progress in a very precise experiment requires adequate progress in the developing of calculational methods and performing the theoretical computations of various observables. This can be consistently done within perturbation theory for processes with large momentum transfer. Nowadays, the standard way to evaluate experimentally measurable quantities from first principles of the theory is to use perturbation methods. Lattice calculations provide an alternative method. The main goal of this paper is to review some of the recent achievements in methods of high order analytical perturbative calculations of a wide class of observable quantities. These quantities are total cross sections, decay widths and structure functions in deep inelastic processes, several key theoretical quantities, such as renormalization group functions, renormalization constants, Wilson coefficient functions, etc. We will present a simplified description of some of the recent calculations. A decisive role in the construction of the SM has been played by experimental studies of so called inclusive processes, in particular deep inelastic lepton-hadron processes like $e^{+}e^{-} \rightarrow \mbox{hadrons}$, deep inelastic $e$, $\mu$ and $\nu$ -scattering, etc. The discovery of scaling of deep inelastic structure functions (Bjorken, 1968, 1969; Yang, 1969) led to the parton model (Feynman, 1969, 1972; see also Drell and Yan, 1971). The explanation of the observed scaling properties has been given by Matveev, Muradyan and Tavkhelidze (1970, 1972), using the universal principle of automodelity and dimensional analysis. The quark counting formulae, allowing one to obtain the high energy asymptotic behavior for cross sections and hadron form factors at large momentum transfers, have been derived by Brodsky and Farrar (1973), and Matveev, Muradian and Tavkhelidze (1973). The discovery of asymptotic freedom in nonabelian gauge field models (Gross and Wilczek, 1973; Politzer, 1973) together with the conception of spin half, fractionally electric charged fundamental constituents of hadrons - quarks (Gell-Mann, 1964; Zweig, 1964) with an additional quantum number color (Bogolyubov, Struminsky and Tavkhelidze, 1965; Tavkhelidze, 1965; Han and Nambu, 1965; also Miyamoto, 1965; Greenberg, 1964), interacting via eight massless, non-abelian, spin 1, self interacting gauge fields - gluons, led to the creation of Quantum Chromodynamics (QCD) (Fritzsch, Gell-Mann and Leutwyler, 1973) - the present theory of strong interactions. For an introductory review of QCD see, e.g., Marciano and Pagels (1978), and a short historical review has been given recently by Tavkhelidze (1994). QCD is based on a local SU$_{c}$(3) symmetry group, which implies the minimal locally gauge invariant Lagrangian density of the model. QCD is a renormalizable quantum field model. There exists well defined rules for removing of ultraviolet divergences from S-matrix amplitudes at each order of the interaction coupling constant. After the renormalization, the calculated physical quantities are free of ultraviolet regularization parameters. The problem of renormalizability of non-abelian gauge theories has been considered since the early 60's (Feynman, 1963; deWitt, 1967; Mandelstam, 1968, etc.). After the Lorentz covariant quantization of gauge fields, based on the path integral approach (Faddeev and Popov, 1967; for a textbook, see also Faddeev and Slavnov, 1980), the proof of renormalizability was given ('t~Hooft, 1971; Lee and Zinn-Justin, 1972, 1973). Besides the short distance effects, in QCD one has to deal with infrared divergences associated with long distance infinities. In other words, in addition to the large parameter (large momentum transferred - $Q^{2}$), there are small parameters such as, for instance, hadron mass - $m$ or momenta of some of the participating particles, and in the calculation one faces senseless large logarithmic contributions $\sim \log m^{2}/Q^{2}$. The infrared divergence problem was considered long ago for QED (Bloch and Nordsieck, 1937; Yennie, Frautschi and Suura, 1961). A modern treatment of this problem in the SM is based on the operator product expansion technique (Wilson, 1969) and factorization theorems. This, in some cases, in particular, for deep inelastic processes, allows one to factorize the large and small distance contributions (Libby and Sterman, 1978; Mueller, 1978; Efremov and Radyushkin, 1980a,b; Radyushkin, 1983; Collins and Soper, 1987; Collins, Soper and Sterman, 1983, 1984, 1985, 1989 and references therein). The concrete prescriptions of dealing with infrared divergent Feynman integrals have been given by Vladimirov (1978, 1980), Pivovarov and Tkachov (1988), Tkachov (1991, 1993), Chetyrkin and Smirnov (1985). For earlier references, see the work by Tkachov (1993). For a textbook, see Collins (1984). The interference of long and short distance effects is still problematic in actual higher order calculations. The group character of renormalizing transformations in quantum field theory was first discovered by Stueckelberg and Peterman (1953), and Gell-Mann and Low (1954) have applied it to study the ultraviolet asymptotics of Green's functions in spinor electrodynamics. The mathematical formalism of the renormalization group has been worked out by Bogolyubov and Shirkov (1955, 1956a,b), and Bogolyubov and Parasyuk (1955a,b, 1956, 1957) have introduced the $R$-operation for subtracting ultraviolet divergent contributions recursively on the level of loop Feynman diagrams. The renormalization group and $R$-operation techniques are the crucial tools in any perturbative calculation within the Standard Model (see the textbook by Bogolyubov and Shirkov, 1980). For a historical review on renormalization group see, for example, Shirkov (1992; also Peterman, 1979) and references therein. For a textbook on the modern renormalization theory and the references, see Collins (1984). The property of asymptotic freedom, the method of renormalization group and factorization theorems are the basis of the present perturbative QCD. In order to relate perturbative QCD to measurable quantities, it was necessary along with the renormalization group, dispersion relations, operator product expansion techniques and factorization, to develop a technique for evaluation of loop Feynman diagrams. Indeed, in each order of perturbation theory, contributions to the physical observables come from a finite set of divergent Feynman integrals with the same number of internal momentum integrations (number of loops). Thus, one has to deal with very singular (ultraviolet and infrared) Feynman integrals, and a correct mathematical apparatus suitable for calculational purposes is necessary. The dimensional regularization technique ('t~Hooft and Veltman, 1972, 1973; Bollini and Giambiagi, 1972; Ashmore, 1972; Cicuta and Montaldi, 1972) for ultraviolet divergent Feynman integrals is based on the idea of integration over the space-time of noninteger dimension less than 4. In this case, the Feynman integrals become well defined, and divergences appear as poles in terms of the deviation from the physical space-time dimension of 4. The important property of dimensional regularization is that it preserves explicit gauge invariance and is very convenient for practical calculations. In fact, almost all recent progress in higher order analytical perturbative calculations has been made within the dimensional regularization, using 't Hooft's (1973) minimal subtraction prescription. The systematic study of strong interaction effects for the various observables in processes with large momentum transfer requires one to evaluate at least the first few coefficients in the perturbative expansion in terms of the strong coupling. Here the problem of calculating multiloop Feynman diagrams arises. The recursive type algorithm for analytical evaluation of one-, two- and three-loop massless, propagator type, dimensionally regularized Feynman diagrams has been given by Chetyrkin and Tkachov (1981), and Tkachov (1981, 1983a). This algorithm, together with the so called Gegenbauer x-space technique (Chetyrkin and Tkachov, 1979) allows one to evaluate an expansion in the Laurent series in $\varepsilon=(4-D)/2$ of all massless propagator type Feynman diagrams up to the three-loop level, where $D$ is the noninteger dimension of the space-time. The above algorithm is applicable to a wide class of problems up to the four-loop level. These are, for instance, calculation of renormalization constants, renormalization group functions, some of the cross sections and decay widths. We note that this algorithm deals only with propagator type massless diagrams. Nevertheless, due to the remarkable properties of dimensional regularization ('t~Hooft and Veltman, 1972) and the minimal subtraction prescription ('t~Hooft, 1973), namely, that the counterterms are polynomials in dimensional parameters within minimal subtraction (Collins, 1974; Speer, 1974; see also the textbook by Collins, 1984), a wide class of problems can be reduced to the evaluation of propagator type diagrams (Vladimirov, 1978, 1980). At high energies, in some cases, it is possible to neglect the masses of participating particles and consider the massless diagrams. The mass corrections of the type $m^{2n}/s^{n}$, where $s$ is the center-of-mass energy squared, can also be evaluated through the calculation of massless diagrams (see, e.g., Gorishny, Kataev and Larin, 1986; Surguladze, 1989a, 1994a,b,c). Feynman graphs can also contain virtual heavy particle propagators regardless of the energy scale of the particular process. If the masses of the virtual particles are much larger than the energy scale, one can neglect them, since their effcts are supressed by powers of large mass, according to the decoupling theorem (Appelquist and Carazzone, 1975). However, in some cases, such effects may not be entirely negligible (Soper and Surguladze, 1994). The prescriptions to study asymptotic expansions of Feynman integrals in powers of $m^{2}/s$ can be obtained from Chetyrkin and Tkachov (1982), Tkachov (1983b,c, 1991, 1993), and Chetyrkin (1991; see also Smirnov 1990, 1991 and references therein). An exact general expression for one-loop, N-point, massive Feynman integrals has been obtained by Davydichev (1991), and Boos and Davydychev (1992). This expression contains the generalized hypergeometric function and is complicated, except for some particular cases. An alternative method for massive Feynman integrals has been suggested by Kotikov (1991). In practice, the calculation of physical quantities within perturbation theory is very cumbersome and tedious already beyond the one-loop level, especially in realistic quntum field theory models, like QCD. However, the recursive type algorithms by Chetyrkin and Tkachov (1981) allow convenient implementation within algebraic programming systems like {\small REDUCE} (Hearn, 1973), {\small SCHOONSCHIP} (Veltman, 1967; Strubbe, 1978; Veltman, 1989) and {\small FORM} (Vermaseren, 1989). Several computer programs were written in the last decade for analytical computation of multiloop Feynman diagrams. Among them we mention the programs which fully implement the above mentioned recursive algorithms. The program {\small LOOPS} (Surguladze and Tkachov, 1989a), written on the {\small REDUCE} system, calculates one- and two-loop massless, propagator type Feynman diagrams for arbitrary structure in the numerator of the integrand and for an arbitrary space-time dimension. The program {\small MINCER} (Gorishny, Larin, Surguladze and Tkachov, 1989), written on the {\small SCHOONSCHIP} system, and the program {\small HEPL}oops (Surguladze, 1992), written on the {\small FORM} system, calculates one-, two- and three-loop massless, propagator type diagrams. The status of the existing program packages has been discussed recently in Surguladze (1994d). The above methods, algorithms and computer programs allow one to make significant progress in high order analytical perturbative calculations of several important physical observables. The other outstanding problem in perturbative calculations is the renormalization group ambiguity of perturbation theory predictions. Indeed, starting from a certain order, the perturbative coefficients become scheme-scale dependent, while it is obvious that the calculated observable cannot depend on any subjective choice of nonphysical parameters. Several approaches have been suggested to deal with the scheme-scale ambiguity problem. Among them we consider the so called {\it fastest apparent convergence} approach (Grunberg, 1980), suggesting one absorb the leading QCD corrections in the definition of the ``effective'' running coupling. We will consider an approach based on the {\it principle of minimal sensitivity} of the physical observables to nonphysical parameters (Stevenson, 1981a,b), and Brodsky, Lepage and Mackenzie (1983) (BLM) method, suggesting one should fix the scale according to the size of the quark vacuum polarization effects. The commensurate scale relations by Brodsky and Lu (1994, 1995) allow one to make scale-fixed perturbative predictions without referring to the particular renormalization prescription. In the recent works, some authors try to predict the perturbative coefficients without calculating the relevant Feynman graphs. First, we mention the method by West (1991) which is based on the renormalizability, analyticity arguments, and the saddle point technique. For comments on this work see Barclay and Maxwell (1992a), Brown and Yaffe (1992), Surguladze and Samuel (1992), and Duncan {\it et al.} (1993). The method of Samuel {\it et al.} (Samuel and Li, 1994a,b,c; Samuel, Li and Steinfelds, 1994a,b,c), based on Pad\`{e} approximants, work surprisingly well for the large number of cases considered. However, a theoretical basis of this method is necessary. Recent developments have put Pad\`{e} approximant method on a more rigorous basis, which justifies its application to perturbation series in QED, QCD, Atomic physics, etc. This is discussed in recent papers (see, e.g., Ellis, Karliner and Samuel, 1995). An alternative method for estimation of higher order perturbative contributions can be obtained based on Stevenson's (1981a,b) approach (Surguladze and Samuel, 1993; Kataev and Starshenko, 1994). The important problem of large-order behavior of perturbation theory has been considered by Barclay and Maxwell (1992b), and Brown and Yaffe (1992). The same problem has been discussed during the past twenty years. The part of papers on the subject have been collected in the book edited by Le Guillou and Zinn-Justin (1990). The application of renormalon calculus in the study of the behaviour of perturbative QCD series is a subject of intensive discussions in the recent literature (see, e.g., Zakharov, 1992; Mueller, 1992; Lovett-Turner and Maxwell, 1994; Vainshtein and Zakharov, 1994, Soper and Surguladze, 1995). After a brief historical review, we turn to the main subject of the present article. Namely, we discuss the analytical high order perturbative calculations of several physical observables which have been completed recently with the help of the above mentioned methods, algorithms and computer programs. First, we consider the analytical calculation of $R(s)$ in electron - positron annihilation at the four-loop level of perturbative QCD (Surguladze and Samuel, 1991a,b; Gorishny, Kataev and Larin, 1991), which turned out to be the most difficult among the problems of this type. This is the first and so far the only four-loop calculation of a physical quantity in QCD. \footnote{ This calculation was attempted earlier by Gorishny, Kataev and Larin (1988) but, unfortunately, errors were found. } As a byproduct, the four-loop $R_{\tau}$ in $\tau$ decay (Gorishny, Kataev and Larin, 1991; Samuel and Surguladze, 1991) and four-loop QED $\beta$ function (Surguladze, 1990; Gorishny, Kataev and Larin, 1990) have been evaluated. \footnote{ For the joint publication of the results of the two independent calculations of the four-loop QED $\beta$-function, see Gorishny, Kataev, Larin and Surguladze (1991a). } For earlier works, we mention, for instance, the calculation of the three-loop correction to $R(s)$ in electron - positron annihilation (Chetyrkin, Kataev and Tkachov, 1979; Dine and Sapirstein, 1979; Celmaster and Gonsalves, 1980), the calculation of the three-loop QCD ${\beta}$ function (Tarasov, Vladimirov and Zharkov, 1980) and the calculation of the three-loop anomalous dimensions of quark masses (Tarasov, 1982). We would also like to list some other three- and two-loop calculations. These are:the calculation of the total decay width of the neutral Higgs boson into hadrons at the three-loop level (Gorishny, Kataev, Larin and Surguladze, 1990, 1991b; Surguladze, 1994a,b), the calculation of the two- and three-loop Wilson coefficients in QCD sum rules (Surguladze and Tkachov, 1986, 1988, 1989b, 1990), the calculation of the two-loop anomalous dimensions of the proton current (Pivovarov and Surguladze, 1991). So far only one five-loop calculation exists. This is the calculation of the five-loop renormalization group functions in $\phi^4$-theory (Kleinert {\it et al.}, 1991). The scope of the present paper is limited and we are not planning to review perturbative QCD. This has already been done and excellent reviews exist. We recommend, for instance, the recent work {\it Handbook of Perturbative QCD} by CTEQ collaboration (Brock, {\it et al.}), edited by G.~Sterman (1993). Here we focus on a somewhat simplified description of the key methods which allow one to perform analytical high order perturbative calculations up to and including the four-loop level. As an example, we will demonstrate the main points of the calculation of $\sigma_{\mbox{\scriptsize{tot}}}(e^{+}e^{-} \rightarrow \mbox{hadrons})$, $\Gamma(\tau^{-} \rightarrow \nu_{\tau} + \mbox{hadrons})$, $\Gamma(H \rightarrow \mbox{hadrons})$ and the QED $\beta$ function. We also outline the calculation of the Wilson coefficient functions of higher twist operators in the operator product expansion and discuss various approaches to resolve the renormalization group ambiguity of perturbation theory predictions. The paper is organised as follows. In the 2nd section we introduce our notation and present some general relations. In this section we discuss the relevant methods and tools of perturbative QCD. We briefly consider the necessary dispersion relation, the operator product expansion (OPE), the renormalization relations and the method for evaluation of the renormalization constants. We also discuss the main ideas of the method of projectors for calculating Wilson coefficients in OPE. In the 3rd section we evaluate the quantity $\Gamma(H \rightarrow \mbox{hadrons})$ at the three-loop level. In the 4th section we calculate the corrections to the correlation functions due to the nonvanishing quark masses. In the 5th section we describe the calculation of the Wilson coefficient functions of the dim=4 operators in the OPE of the two-point correlation function of quark currents. In the 6th section we describe the four-loop calculation of $\sigma_{\mbox{\scriptsize{tot}}}(e^{+}e^{-} \rightarrow \mbox{hadrons})$. Sections 7 and 8 are dedicated to the evaluation of $\Gamma(\tau^{-} \rightarrow \nu_{\tau} + \mbox{hadrons})$ and the QED $\beta$ function respectively. In section 9 we discuss the problem of the renormalization group ambiguity of perturbative QCD results. As an example, we consider calculated quantities and use the known approaches to try to fix the scheme-scale parameter within the one parametric family of MS-type schemes. Next, we outline the original method of scheme-invariant analysis and optimization procedure by Stevenson (1981a,b). The paper ends with summarizing notes. \renewcommand{\thesection}{\Roman{section}} \section{\bf Calculational methods} \renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0} \subsection{\tenbf Notation and general relations of perturbative QCD} Throughout this paper we work within the standard model of strong interactions - QCD. For a review on QCD see, for example, Marciano and Pagels (1978), Mueller (1981), Reya (1981), and Altarelli (1982). For a textbook see, e.g., Yndurain (1983), Quigg (1986), Muta (1987), and Ellis and Stirling (1990). For the most recent source see, e.g., {\it Handbook of Perturbative QCD} by CTEQ collaboration (Brock {\it et al.}), edited by G.~Sterman (1993). The four-loop QED calculations will be discussed in section 8. The Lagrangian density of standard QCD is \begin{eqnarray} L(x)=-1/4(G_{\mu\nu}^{a})^{2}-\frac{1}{2\alpha_G} (\partial^{\mu}A_{\mu}^{a})^{2} +\sum_{f}\overline{q}_{f}(i\hat{\partial}-m_{f})q_{f} +g\sum_{f}\overline{q}_{f}T^{a}\hat{A}^{a}q_{f} \nonumber\\ +\partial^{\mu}c^{a^{\dag}}(\partial_{\mu}\delta^{ac} +gf^{abc}A_{\mu}^{b})c^{c}, \label{eq:lagrangian} \end{eqnarray} where $G_{\mu\nu}^{a}=\partial_{\mu}A_{\nu}^{a}-\partial_{\nu}A_{\mu}^{a} +gf^{abc}A_{\mu}^{b}A_{\nu}^{c}$ \ ($a=1,2,...,8$) are the Yang-Mills field (Yang and Mills, 1954) strengths, \ $A^{a}$ and $q_{f}$ are gluon and quark fields, $m_{f}$ are the quark masses, \ $c^{a}$ are the Faddeev-Popov ghosts and $\alpha_{\mbox{\tiny G}}$ is the gauge parameter. We use the standard notation $\hat{\partial}=\gamma^{\mu}\partial_{\mu}$ and $\hat{A}^{a}=\gamma^{\mu}A_{\mu}^{a}$. The index $f$ enumerates the quark flavors, total number of which is $N$. The generators $T^{a}$ of the SU$_{\mbox{\scriptsize c}}$(N) gauge group, the structure constants $f^{abc}$ and $d^{abc}$ obey the following relations \begin{eqnarray} [T^a,T^b]=if^{abc}T^c,\hspace{2mm} \{T^a,T^b\}=\frac{1}{N}\delta^{ab}+d^{abc}T^c, \nonumber\\ f^{acd}f^{bcd}=C_A\delta^{ab},\hspace{2mm} T^aT^a=C_F\hat{\bf 1}, \hspace{2mm} \mbox{tr}T^aT^b=T\delta^{ab}. \label{eq:casimirs} \end{eqnarray} The eigenvalues of the Casimir operators for the adjoint ($N_A=8$) and the fundamental ($N_F=3$) representations of SU$_{\mbox{\scriptsize c}}$(3) are \begin{equation} C_{A}=3,\hspace{2mm}C_{F}=4/3, \hspace{2mm} \mbox{and} \hspace{2mm} T=1/2,\hspace{2mm} d^{abc}d^{abc}=40/3. \label{eq:casimirsnum} \end{equation} We use the standard QCD Feynman rules (see, e.g., Abers and Lee, 1973; Muta, 1987). \vspace{3mm} {\bf Propagators} \vspace{4mm} \hspace{7mm} quark \hspace{43mm} $= \frac{1}{i}\frac{m+\hat{P}} {m^2-P^2}\delta_{ij}$ \vspace{13mm} \hspace{7mm} gluon \hspace{43mm} $= \frac{1}{i}\frac{\delta_{ab}}{P^2} \left[g^{\mu\nu} -(1-\alpha_{\mbox{\tiny G}}) \frac{P_{\mu}P_{\nu}}{P^2}\right]$ \vspace{13mm} \hspace{7mm} ghost \hspace{43mm} $= \frac{1}{i}\frac{\delta_{ab}}{P^2}$ \vspace{10mm} {\bf Vertices} \vspace{7mm} \hspace{7mm} quark-quark-gluon \hspace{43mm} $= ig\gamma^{\mu}T^{a}_{ij}$ \vspace{29mm} \hspace{7mm} ghost-ghost-gluon \hspace{43mm} $= igf^{abc}P_{\mu}$ \vspace{26mm} \hspace{7mm} 3-gluon \hspace{43mm} $=gf^{abc}[g_{\mu\nu}(q-p)_{\lambda} +g_{\nu\lambda}(k-q)_{\mu} +g_{\mu\lambda}(p-k)_{\nu}]$ \vspace{29mm} \hspace{7mm} 4-gluon \hspace{43mm} $=ig^2[f^{abe}f^{cde}(g^{\mu\lambda} g^{\nu\rho} -g^{\mu\rho}g^{\nu\lambda})$ \hspace{74mm} $+f^{ace}f^{bde}(g^{\mu\nu} g^{\lambda\rho}-g^{\mu\rho}g^{\nu\lambda})$ \hspace{74mm} $+f^{ade}f^{cbe}(g^{\mu\lambda}g^{\nu\rho} -g^{\mu\nu}g^{\lambda\rho})]$ \vspace{11mm} The sum of all momenta coming in each vertex of the Feynman diagram is zero (momentum conservation). \vspace{4mm} {\bf Factors} \vspace{3mm} (-1) for each closed fermion or ghost loop \vspace{3mm} Statistical factors (for derivations see, e.g., 't~Hooft and Veltman, 1973): \vspace{9mm} $\frac{1}{2}$ for each graph (subgraph) \vspace{14mm} $\frac{1}{6}$ for each graph (subgraph) \hspace{5cm} etc. \vspace{9mm} {\bf Integration} \vspace{3mm} Each loop corresponds to the integration \hspace{7mm} $\int\frac{d^4P}{(2\pi)^4}$. \vspace{7mm} In general, the Feynman integral constructed according to the above rules is divergent. There are two kind of divergences. One, the so called ultraviolet (UV) divergence is due to large integration momenta and the other one - the so called infrared divergence is associated with the small integration momenta in the massless limit. The most convenient regularization of Feynman integrals is dimensional regularization ('t~Hooft and Veltman, 1972; Bollini and Giambiagi, 1972; Ashmore, 1972; Cicuta and Montaldi, 1972), where the space-time dimension is analytically continued from the physical value, 4, to a complex value $D=4-2\varepsilon$. In the limit $\varepsilon \rightarrow 0$, the divergences appear as poles $1/\varepsilon$, defining the counterterms. One of the remarkable properties of dimensional regularization is that the Ward identities implied by gauge invariance are maintained for arbitrary space-time dimension D, in contrast with the old Pauli-Villars regularization (Pauli and Villars, 1949). Another useful property is a convenience in practical multiloop calculations. Thus, in dimensional regularization we formally replace $\int\frac{d^4P}{(2\pi)^4}$ $\rightarrow$ $\int\frac{d^{D}P}{(2\pi)^D}$. It is straightforward to extend the all necessary tensor algebra into $D$-dimensions. For example, $g^{\mu\nu}g_{\mu\nu}=D$, $\mbox{Tr}{\gamma_{\mu}\gamma_{\nu}} =2^{D/2}g_{\mu\nu}$, etc. For the complete list of formulae see, e.g., Collins (1984) and also Narison (1982). Note, however, that the extension of the usual definition of the matrix $\gamma_{5}$\\ \begin{center} $\gamma_{5} = \frac{1}{4!}\varepsilon_{\alpha\beta\mu\nu}\gamma_{\alpha} \gamma_{\beta}\gamma_{\mu}\gamma_{\nu}$\\ \end{center} is not straightforward. The totally antisymmetric tensor $\varepsilon_{\alpha\beta\mu\nu}$ is defined only in the four-dimensional space. In some cases the calculation of the quantities involving $\gamma_{5}$ is still possible within dimensional regularization. For a discussion of the problem of $\gamma_{5}$ in dimensional regularization see Delbourgo and Akyeampong (1974), Trueman (1979), Bonneau (1980), Narison (1982), Collins (1984), and Larin (1993). For a calculation involving $\gamma_{5}$ within dimensional regularization see, e.g., Pivovarov and Surguladze (1991). In order to get finite physical quantities, the divergences in dimensionally regularized Feynman integrals, appearing as poles in $1/\varepsilon$, need to be subtracted by adopting of some specific rule. This rule is usually called a renormalization scheme. Throughout this paper we use 't~Hooft's minimal subtraction (MS) type scheme ('t~Hooft, 1973). The subtraction of divergences is equivalent to the redefinition (renormalization) of the parameters (coupling, mass and gauge fixing parameter) and fields in the original ``bare'' lagrangian \begin{center} $\alpha_s^{\mbox{\tiny B}}=\mu^{2\varepsilon}Z_{\alpha_s}\alpha_s$, \hspace{1cm} $(g^{2}/4\pi\equiv\alpha_{s})$ \end{center} \begin{equation} m^{\mbox{\tiny B}} = mZ_{m}, \label{Zdefinitions} \end{equation} \begin{center} $\alpha_{\mbox{\tiny G}}^{\mbox{\tiny B}} = \alpha_{\mbox{\tiny G}} Z_{\mbox{\tiny G}}$. \end{center} $\mu$ is a quantity of dimension of mass which is introduced within dimensional regularization in order to make an action dimensionless. Superscript ``B'' denote the unrenormalized quantity. We renormalize the gluon, quark and ghost fields analogously. Within the MS scheme the N-point Green function is renormalized in the following way \begin{equation} \Gamma(p_1,...,p_N,g,m,\alpha_{\mbox{\tiny G}},\mu) = Z_{\Gamma}\Gamma^{\mbox{\scriptsize B}} (p_1,...,p_N,g,m,\alpha_{\mbox{\tiny G}}), \label{Zgreen} \end{equation} where $Z_{\Gamma}$ is a polynomial in $1/\varepsilon$, and thus multiplying by $Z_{\Gamma}$, we subtract only pole parts from the divergent $\Gamma^{\mbox{\scriptsize B}}$. The evaluation of the renormalization constants $Z$ will be discussed in the next subsections. It is easy to see that the $\mu$ parameter entered through the renormalization and hence the unrenormalized Green's function is independent of $\mu$ \begin{displaymath} \mu\frac{d}{d\mu}\Gamma^{\mbox{\scriptsize B}}(p_1,...,p_N,g,m, \alpha_{\mbox{\tiny G}}) = 0. \end{displaymath} Using eq.\ (\ref{Zgreen}) and expanding the full derivative we get the renormalization group equation in the following form \begin{equation} \left[\mu^2\frac{\partial}{\partial\mu^2} +\beta(\alpha_s)\alpha_s\frac{\partial}{\partial\alpha_s} -\gamma_{m}(\alpha_s)m\frac{\partial}{\partial m} +\beta_{\mbox{\tiny G}}(\alpha_s)\frac{\partial} {\partial\alpha_{\mbox{\tiny G}}} -\gamma_{\Gamma}\right] \Gamma(p_1,...,p_N,m,\alpha_s,\alpha_{\mbox{\tiny G}},\mu) = 0. \label{RGE} \end{equation} The QCD renormalization group functions - the $\beta$-function and the anomalous dimension functions - $\gamma$ are defined in the following way \begin{displaymath} \alpha_s\beta(\alpha_s)=\mu^2\frac{d\alpha_s}{d\mu^2}, \end{displaymath} \begin{displaymath} \beta_{\mbox{\tiny G}}(\alpha_s) =\mu^2\frac{d\alpha_{\mbox{\tiny G}}}{d\mu^2}, \end {displaymath} \begin{equation} \gamma_{m}(\alpha_s)=-\frac{\mu^2}{m}\frac{dm}{d\mu^2}, \label{eq:RGfunctions} \end{equation} \begin{displaymath} \gamma_{\Gamma}(\alpha_s)=\frac{\mu^2}{Z_{\Gamma}}\frac{dZ_{\Gamma}}{d\mu^2}, \end{displaymath} with bare coupling and mass fixed. In the present paper we use the renormalization group equation in the above form. The other forms are also known in the literature. The group properties of the renormalization was first discovered by Stueckelberg and Peterman (1953). The ultraviolet asymptotics of the Green function was studied by Gell-Mann and Low (1954) in quantum electrodynamics using the group of multiplicative renormalizations. The renormalization group formalism was further developed in the original works by Bogolyubov and Shirkov (1955, 1956a,b). For the detailed monograph see Bogolyubov and Shirkov (1980). The renormalization group equation was studied by Ovsyannikov (1956), Callan (1970), and Symanzik (1970). For a recent historical review see Shirkov (1992) and references therein. The renormalization group $\beta$-function and anomalous dimensions of quark masses are calculated up to the three-loop level (Tarasov, Vladimirov and Zharkov, 1980; Tarasov, 1982). The QCD $\beta$-function up to and including the three-loop level in MS type schemes is \begin{equation} \beta(\alpha_{s})=-\beta_{0}\frac{\alpha_s}{\pi} -\beta_1(\frac{\alpha_s}{\pi})^2 -\beta_2(\frac{\alpha_s}{\pi})^3+O(\alpha_s^4), \label{eq:beta} \end{equation} where (Tarasov, Vladimirov and Zharkov, 1980) \vspace{3mm} \noindent $\beta_0=\frac{1}{4}\biggl(\frac{11}{3}C_A -\frac{4}{3}TN\biggr)$, \noindent $\beta_1=\frac{1}{16}\biggl(\frac{34}{3}C_A^2-\frac{20}{3}C_ATN -4C_FTN\biggr)$, \noindent $\beta_2=\frac{1}{64}\biggl( \frac{2857}{54}C_A^3-\frac{1415}{27}C_A^2TN +\frac{158}{27}C_AT^2N^2-\frac{205}{9}C_AC_FTN +\frac{44}{9}C_FT^2N^2+2C_F^2TN\biggr)$. \vspace{3mm} \noindent The quark mass anomalous dimension up to and including three-loop level is \begin{equation} \gamma_{m}(\alpha_{s})=\gamma_{0}\frac{\alpha_s}{\pi} +\gamma_1(\frac{\alpha_s}{\pi})^2 +\gamma_2(\frac{\alpha_s}{\pi})^3+O(\alpha_s^4), \label{eq:gamma} \end{equation} where (Tarasov, 1982) \vspace{3mm} \noindent $\gamma_0=\frac{3}{4}C_F$, \noindent $\gamma_1=\frac{1}{16}\biggl(\frac{3}{2}C_{F}^2+\frac{97}{6}C_{F}C_A -\frac{10}{3}C_FTN\biggr)$, \noindent $\gamma_2=\frac{1}{64}\biggl[\frac{129}{2}C_{F}^3-\frac{129}{4}C_{F}^{2}C_{A} +\frac{11413}{108}C_{F}C_{A}^2-(46-48\zeta(3))C_{F}^{2}TN$ \hspace{7cm} $-\biggl(\frac{556}{27}+48\zeta(3)\biggr)C_{F}C_{A}TN -\frac{140}{27}C_{F}T^2N^2\biggr]$. \vspace{3mm} As it was shown by Caswell and Wilczek (1974) and Banyai, Marculescu and Vescan (1974), the above renormalization group functions are gauge independent, which greatly simplifies their evaluation. In fact, the QCD $\beta$-function and the quark mass anomalous dimension have been evaluated in the Feynman gauge $\alpha_{\mbox{\tiny G}}=1$. We note that the perturbative coefficients of the renormalization group functions are the same within the one parametric family of the MS type schemes. Note also the independence of these perturbative coefficients on the quark masses by their definition within the MS type schemes. \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf Vacuum polarization function and Dispersion relation} \indent The vacuum polarization functions for various types of quark currents are crucial in the theoretical evaluation of total cross sections and decay widths. Indeed, for example, the quantity $\sigma_{\mbox{\scriptsize{tot}}}(e^{+}e^{-}\rightarrow \mbox{hadrons})$, according to the well known optical theorem (see, e.g., the textbook by Bogolyubov and Shirkov, 1980), is proportional to the imaginary part of the function $\Pi(-q^2+i0)$, defined from the hadronic vacuum polarization function \begin{equation} \Pi_{\mu\nu}(q)=i\int e^{iqx}<Tj_\mu(x)j_\nu(0)>_0d^4x =(g_{\mu\nu}Q^2-Q_{\mu}Q_{\nu})\Pi(Q^2)\frac{1}{(4\pi)^2}. \label{eq:pifunction} \end{equation} Here, $j_{\mu}(x)=Q_f\overline{q}_f\gamma_{\mu}q_f$, $Q_f$ is the electric charge of the quark of flavor $f$ and $Q^2=-q^2$ is the Euclidean momentum squared. The sum over all participating quark flavors is assumed in $\Pi$. The transverse form in the r.h.s. is conditioned by the conservation of electromagnetic currents. In this paper we also consider the two-point function of quark axial vector currents associated with the quantity $\Gamma(Z \rightarrow \mbox{hadrons})$ and two-point function of quark scalar currents associated with the quantity $\Gamma(H \rightarrow \mbox{hadrons})$ - the total decay width of the neutral Standard Model Higgs boson into hadrons. The renormalized vacuum polarization function obeys the dispersion relation \begin{equation} \Pi(Q^2) = \frac{4}{3}\int_{s_{0}}^{\infty}\frac{R(s)}{s+Q^2}ds - \mbox{subtractions}, \label{eq:disprelat} \end{equation} where \begin{equation} R(s)=\frac{\sigma_{\mbox{\scriptsize{tot}}}(e^{+}e^{-}\rightarrow \mbox{hadrons})} {\sigma(e^{+}e^{-}\rightarrow \mu^{+}\mu^{-})} =\frac{3}{4\pi} \mbox{Im}\Pi(s+i0). \label{eq:Rsdefin} \end{equation} Recall also that the muon pair production cross-section $\sigma(e^{+}e^{-}\rightarrow \mu^{+}\mu^{-})=4\pi\alpha^2/3s$, where $\alpha=e^2/4\pi$ is the electromagnetic fine structure constant. The above dispersion relation allows one to connect the experimentally measurable quantity $R(s)$ to the $\Pi(Q^2)$ calculable perturbatively in the deep Euclidean region ($Q^2$ is large compared to the typical hadron mass). For the discussion on theoretical calculability of R(s) see earlier references: Adler (1974), Appelquist and Politzer (1975), De~R\'{u}jula and Georgi (1976), Poggio, Quinn and Weinberg (1976), Shankar (1977), and Barnett, Dine and McLerran (1980). The combination of the idea of local duality in the dispersion relations (Logunov, Soloviov and Tavkhelidze, 1967) and the Operator Product Expansion technique (Wilson, 1969) became a basis of various versions of QCD sum rules (Shifman, Vainshtein and Zakharov, 1979; Novikov {\it et al.}, 1978, 1985; Krasnikov, Pivovarov and Tavkhelidze, 1983; Shifman, 1992 and references therein). The methods of QCD sum rules are widely used to obtain quantitative information on the observed hadron spectrum and to extract the fundamental theoretical parameters. In practice, sometimes it is more convenient to introduce the Adler function (Adler, 1974) \begin{equation} D(Q^2)=-\frac{3}{4}\frac{\partial}{\partial \log Q^2}\Pi(Q^2) =Q^2\int_{s_0}^{\infty}\frac{R(s)}{(s+Q^2)^2}ds. \label{eq:Ddefin} \end{equation} Derivative here avoids an inconvenient extra subtraction in the r.h.s. The leading (parton) approximation of $D(Q^2)$ in the zero quark mass limit coincides with $R(s)$ \begin{equation} D(Q^2) = 3\sum_{f}Q_{f}^{2}, \label{eq:D0} \end{equation} where the sum runs over all participating quark charges at the given energy. 3 stands for the number of different colors. The leading ``non-QCD'' contribution is completely free of ultraviolet divergences, while the $\Pi(Q^2)$ needs an additive renormalization even at the leading order. At higher orders of perturbative expansion of the $D$-function the ultraviolet divergences appear and one should employ a procedure (usually called renormalization scheme) for their subtraction order-by-order. Because of ambiguity in the choice of subtraction scheme, the amplitude calculated within the perturbation theory depends on nonphysical parameters. Within the one-parametric family of the MS type schemes (t~'Hooft, 1973) such a parameter is usually called $\mu$. Thus, up to power corrections, the $D$-amplitude will be a function of $\log(\mu^2/Q^2)$ and the strong coupling $\alpha_s$. On the other hand, since $D$ is connected to the observable $R(s)$, it can not depend on our subjective choice of nonphysical parameter $\mu$. This can be achieved if the strong coupling becomes a function of $\mu$, providing independence of observables on the choice of parameter $\mu$. Here, it is asumed that all orders of perturbation theory are summed up. Otherwise, if one considers a truncated series, the $\mu$ dependence remains. The problem of scheme-scale dependence and some possible solutions will be discussed later in this review. The set of transformations which leave observables independent of renormalization parameters has a group character and forms the renormalization group. The renormalization group in renormalizable theories (like QCD) fixes the dependence of the coupling on the $\mu$-parameter. The function $D(Q^2)$ calculated in perturbative QCD within the MS type schemes obeys the renormalization group equation \begin{equation} \left(\mu^2\frac{\partial}{\partial\mu^2} +\beta(\alpha_s)\alpha_s\frac{\partial}{\partial\alpha_s} -\gamma_m(\alpha_s)m\frac{\partial} {\partial m}\right)D(\mu^2/Q^2,m,\alpha_s)=0. \label{eq:RGD} \end{equation} Below we consider the limit of the massless light quarks and the infinitely large top mass which decouples (Appelquist and Carazzone, 1975). The solution of eq.\ (\ref{eq:RGD}) at $\mu^2=Q^2$ is \begin{equation} D(\mu^2/Q^2,\alpha_s(\mu))=D(1,\alpha_{s}(Q))=\sum_{i\geq0} R_i(\alpha_{s}(Q)/\pi)^i, \label{eq:RGDsolut} \end{equation} where the $\alpha_{s}(\mu^2)$ is the running coupling, usually parametrized up to the three-loop level as follows \begin{equation} \frac{\alpha_s(\mu^2)}{\pi}=\frac{1}{\beta_0 L}-\frac{\beta_1 \log L} {\beta_0^3 L^2}+\frac{1}{\beta_0^5 L^3}(\beta_1^2 \log^2 L-\beta_1^2 \log L +\beta_2 \beta_0-\beta_{1}^{2})+O(L^{-4}), \label{eq:Asparametr} \end{equation} where $L=\log (\mu^2/\Lambda^2)$. Parametrization (\ref{eq:Asparametr}) has the same form and the QCD $\beta$-function coefficients are the same within the MS type schemes. The scale parameter $\Lambda$ depends on the particular modification of the MS prescription. In fact, $\Lambda$ is used to parametrize other versions of renormalization prescription as well. It is shown by Celmaster and Gonsalves (1979) that the transformation relations valid to all orders between $\Lambda$'s defined by any two renormalization prescription can be deduced from a one-loop calculation. Comparing the bare coupling constants within different renormalization prescriptions and using the results for the one-loop renormalization constants and the property of asymptotic freedom, one obtains for, e.g., momentum subtraction (MOM) and MS schemes (Celmaster and Gonsalves, 1979) \begin{equation} \Lambda_{\mbox{\scriptsize MS}}=\Lambda_{\mbox{\scriptsize MOM}} \mbox{exp}\biggl[\frac{A(\alpha_{\mbox{\tiny G}},N)} {4\beta_0}\biggr], \label{eq:MOMMS} \end{equation} where \begin{eqnarray} \lefteqn{\hspace{-75mm} A(\alpha_{\mbox{\tiny G}},N)=C_A\biggl[ -\frac{11}{6}(\gamma_{\mbox{\tiny E}}-\ln 4\pi) +\frac{11}{3}+\frac{23}{72}I +\frac{3}{8}\alpha_{\mbox{\tiny G}}(1-I) -\frac{1}{12}\alpha_{\mbox{\tiny G}}^2(3-I) +\frac{1}{24}\alpha_{\mbox{\tiny G}}^3\biggr]} \nonumber\\ && \hspace{-64mm} +TN\biggl[\frac{2}{3}(\gamma_{\mbox{\tiny E}}-\ln 4\pi) -\frac{4}{3}-\frac{8}{9}I\biggr] \label{eq:A} \end{eqnarray} and the integral \begin{equation} I=-2\int_0^1 \frac{\ln x}{x^2-x+1}dx=2.3439072... \label{eq:I} \end{equation} One note due to Stevenson (1981b, 1994) is in order. Despite its convenient form, the parametrization (\ref{eq:Asparametr}) produces an additional ambiguity due to the freedom with a particular definition of $\Lambda$ parameter, even when the renormalization prescription is already specified. This problem was discussed by Abbot (1980), Shirkov (1980), Stevenson (1981b), Monsay and Rosenzweig (1981) and Radyushkin (1983). In fact, one can take advantage of this freedom in the choice of $\Lambda$ and try to optimize the expansion in $1/L$. Indeed, as was shown by Radyushkin (1983), if one takes $0.6\Lambda$ in eq.\ (\ref{eq:Asparametr}) instead of standard (Buras, Floratos, Ross and Sachrajda, 1977) $\Lambda$ then the $1/L^2$ and $1/L^3$ terms contribute only a few percent for a reasonably wide range of $\mu$. On the other hand, Stevenson (1981b, 1994) has suggested to avoid the entire problem of ambiguity in the definition of $\Lambda$ by abandoning the $1/L$ expansion and solving the renormalization group equation (\ref{eq:RGfunctions}) for $\alpha_s$ and resulting transcendental equation numerically, using the truncated $\beta$ function. According to the operator product expansion technique (Wilson, 1969), one can separate perturbative and nonperturbative contributions to the function $\Pi(Q^2)$. As shown by Shifman, Vainshtein and Zakharov (1979), this function can be represented in the following form \begin{equation} \Pi(Q^2)=\mbox{perturbation theory} +\sum_{n\geq2}\frac{C_{n}(Q)<O_n>_0}{Q^{2n}} +\mbox{instanton contributions,} \label{eq:PIOPE} \end{equation} where $<O_n>_{0}$ denote vacuum condensates parametrizing the nonperturbative contributions and $C_n(Q)$ are their coefficient functions. The last term in the above equation describes the instanton contributions, which, in the case of electromagnetic currents, was estimated to be small (Krasnikov and Tavkhelidze, 1982; Kartvelishvili and Margvelashvili, 1995). The coefficient functions of the condensates can be calculated within perturbation theory. High order perturbative corrections to the coefficient functions of dimension 4 and 6 power terms have been calculated in Loladze, Surguladze and Tkachov (1984, 1985), Surguladze and Tkachov (1989b, 1990), Chetyrkin, Gorishny and Spiridonov (1985), and Lanin, Chetyrkin and Spiridonov (1986). In subsection E we discuss the method for evaluation of Wilson coefficient functions. Examples will be outlined in section 4. Note, that we consider the region of very high energies where, in fact, only perturbation theory contributions survive in $\Pi(Q^2)$. The nonperturbative corrections could have some (small) effect in the case of, for instance, $\tau$ lepton decay (see section 7). Note also that, in fact, the effects of neglected light quark masses are not entirely negligible in some phenomenological applications (see section 4). \vspace{2cm} \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf Renormalization relations} \indent There are several approaches for the ultraviolet renormalization of Green's functions known in the literature. Throughout this paper we use 't Hooft's minimal subtruction method ('t Hooft, 1971, 1973). For alternative prescriptions we refer to the works by Weinberg (1967), Gell-Mann and Low (1954), Callan (1970), Symanzik (1970), and Collins, Wilczek and Zee (1978). For an analysis of various renormalization methods see Collins and Macfarlane (1974). For a review see, e.g., Narison (1982), the textbook by Collins (1984) and references therein. We focus on the renormalization relations for the two point correlation function of quark currents relevant for the further evaluation of total cross sections and decay widths. It is known that the vacuum polarization function is renormalized additively \begin{equation} \Pi(\mu^2/Q^2,\alpha_s) =\Pi^{\mbox{\scriptsize B}} (\mu^2/Q^2,\alpha_s^{\mbox{\tiny B}})+Z_{\Pi}\equiv \mbox{finite}. \label{eq:PiR} \end{equation} The bare coupling $\alpha_s^{\mbox{\tiny B}}$ is related to the renormalized one by the relation (\ref{Zdefinitions}). The perturbative expansion for $Z_{\alpha_s}$ can be found based on eqs.\ (\ref{eq:RGfunctions}) and (\ref{eq:beta}), the MS definition of $Z_{\alpha_s}$ and the renormalization group equation \begin{equation} \mu^2\frac{d}{d\mu^2}\alpha_s^{\mbox{\tiny B}}=0. \label{eq:AsR1} \end{equation} We obtain \begin{equation} Z_{\alpha_s}=1-\frac{\alpha_s}{\pi}\frac{\beta_0}{\varepsilon} +\left(\frac{\alpha_s}{\pi}\right)^2 \left(\frac{\beta_0^2}{\varepsilon^2} -\frac{\beta_1}{2\varepsilon}\right) -\left(\frac{\alpha_s}{\pi}\right)^3 \biggl(\frac{\beta_0^3}{\varepsilon^3} -\frac{7}{6}\frac{\beta_0\beta_1}{\varepsilon^2} +\frac{\beta_2}{3\varepsilon}\biggr) +O(\alpha_s^4). \label{eq:AsR} \end{equation} In general, the polarization function depends on quark masses and we will need the relation between ``bare'' and renormalized masses up to $O(\alpha_s^2)$ (Tarasov, 1982) \begin{eqnarray} \lefteqn{(m_{f}^{\mbox{\tiny B}})^2 = m_{f}^{2} \biggl\{ 1 -\left(\frac{\alpha_{s}}{4\pi}\right)\frac{6C_F}{\varepsilon} +\left(\frac{\alpha_{s}}{4\pi}\right)^2C_{F} \biggl[ \biggl(11C_{A}+18C_{F}-4TN\biggr)\frac{1}{\varepsilon^2}} \nonumber\\ && \quad \hspace{43mm} -\biggl(\frac{97}{6}C_{A}+\frac{3}{2}C_{F}-\frac{10}{3}TN\biggr) \frac{1}{\varepsilon} \biggr]+O(\alpha_{s}^{3}) \biggr\}. \label{eq:mrenorm} \end{eqnarray} Within the minimal subtraction prescription ('t Hooft, 1973) the renormalization constant $Z_{\Pi}$ can be expressed as the following double sum \begin{equation} Z_{\Pi}=\sum_{\stackrel{-l\leq k<0}{l>0}} \left(\frac{\alpha_s}{\pi}\right)^{l-1}Z_{l,k} \varepsilon^k, \label{eq:Zexpans} \end{equation} where $Z_{lk}$ are numbers. Furthermore, for the ``bare'' vacuum polarization function one has the following expansion in a perturbation series \begin{equation} \Pi^{\mbox{\scriptsize B}}\left(\frac{\mu^2}{Q^2}, \alpha_s^{\mbox{\tiny B}}\right)= \sum_{\stackrel{-l\leq k}{l>0}} \left(\frac{\alpha_s^{\mbox{\tiny B}}}{\pi}\right)^{l-1} \left(\frac{\mu^2}{Q^2}\right)^{l\varepsilon}\Pi_{l,k}\varepsilon^k, \label{eq:Piexpans} \end{equation} where the first index denotes the number of loops of the corresponding Feynman diagrams at the given order of $\alpha_s$. Substituting eqs.\ (\ref{eq:Piexpans}) and (\ref{eq:PiR}) into the definition (\ref{eq:Ddefin}), after the renormalization of the coupling via (\ref{eq:AsR}) we obtain at $\mu^2=Q^2$ \begin{eqnarray} \lefteqn{\hspace{-7mm}D(\alpha_{s})=\frac{3}{4} \biggl\{ \Pi_{1,-1} +\frac{\alpha_{s}}{\pi} \biggl[2\Pi_{2,-2}\frac{1}{\varepsilon}+2\Pi_{2,-1}\biggr] +\left(\frac{\alpha_{s}}{\pi}\right)^2 \biggl[ \frac{1}{\varepsilon^2}\left(3\Pi_{3,-3}-2\beta_0\Pi_{2,-2}\right)} \nonumber\\ && \quad \hspace{51mm} +\frac{1}{\varepsilon} \left(3\Pi_{3,-2}-2\beta_0\Pi_{2,-1}\right) +\left(3\Pi_{3,-1}-2\beta_0\Pi_{2,0}\right)\biggr] \nonumber\\ && \quad +\left(\frac{\alpha_{s}}{\pi}\right)^3 \biggl[ \frac{1}{\varepsilon^3}\left(4\Pi_{4,-4}-6\beta_0\Pi_{3,-3} +2\beta_0^2\Pi_{2,-2}\right) \nonumber\\ && \quad \hspace{16mm} +\frac{1}{\varepsilon^2}\left(4\Pi_{4,-3}-6\beta_0\Pi_{3,-2} -\beta_1\Pi_{2,-2}+2\beta_0^2\Pi_{2,-1}\right) \nonumber\\ && \quad \hspace{16mm} +\frac{1}{\varepsilon}\left(4\Pi_{4,-2}-6\beta_0\Pi_{3,-1} -\beta_1\Pi_{2,-1}+2\beta_0^2\Pi_{2,0}\right) \nonumber\\ && \quad \hspace{16mm} +\left(4\Pi_{4,-1}-6\beta_0\Pi_{3,0} -\beta_1\Pi_{2,0}+2\beta_0^2\Pi_{2,1}\right) \biggr] +O(\alpha_{s}^{4}) \biggr\}. \label{eq:DDexpans} \end{eqnarray} Because of the renormalization group invariance of $D(\mu^2/Q^2,\alpha_s)$, in the above equation we take $\mu^2=Q^2$ to avoid unnecessary logarithms. The renormalized expression for the $D$-function must be finite in the limit $\varepsilon \rightarrow 0$. Thus the coefficients of pole terms must vanish identically. This implies relations between the perturbative coefficients of $\Pi$ and the QCD $\beta$-function. First, we note that prior to any renormalization the leading poles must cancel at each order of $\alpha_s$ in the sum of all relevant Feynman diagrams. As shown by the actual calculation, this happens in each gauge invariant set of diagrams. \begin{equation} \Pi_{4,-4} = \Pi_{3,-3} = \Pi_{2,-2} = 0. \label{eq:Pirelat0} \end{equation} Moreover, from the cancellation of nonleading poles we get \begin{displaymath} 3\Pi_{3,-2}-2\beta_{0}\Pi_{2,-1} = 0, \end{displaymath} \begin{equation} 4\Pi_{4,-3}-6\beta_{0}\Pi_{3,-2}+2\beta_{0}^2\Pi_{2,-1} = 0, \label{eq:Pirelat1} \end{equation} \begin{displaymath} 4\Pi_{4,-2}-6\beta_{0}\Pi_{3,-1}-\beta_{1}\Pi_{2,-1} +2\beta_{0}^2\Pi_{2,0} = 0. \end{displaymath} The above relations provide powerful tests of the calculation at its intermediate stages and are crucial. {}From eq.\ (\ref{eq:PiR}) we see that fully renormalized $\Pi(Q^2,\alpha_s)$ must be finite. Thus, substituting eqs.\ (\ref{eq:AsR}) - (\ref{eq:Piexpans}) and (\ref{eq:Pirelat0}) in eq.\ (\ref{eq:PiR}) we obtain the following expression for the divergent part of $\Pi(\mu^2/Q^2,\alpha_s)$ at $\mu^2=Q^2$ \begin{eqnarray} \lefteqn{\mbox{div}\Pi(\alpha_{s}) = \frac{1}{\varepsilon}(\Pi_{1,-1}+Z_{1,-1}) + \frac{\alpha_{s}}{\pi} \biggl[ \frac{1}{\varepsilon}(\Pi_{2,-1}+Z_{2,-1})\biggr]} \nonumber \\ && \quad +\left(\frac{\alpha_{s}}{\pi}\right)^2 \biggl[ \frac{1}{\varepsilon^2}(\Pi_{3,-2} -\beta_0\Pi_{2,-1}+Z_{3,-2}) + \frac{1}{\varepsilon}(\Pi_{3,-1} - \beta_0\Pi_{2,0}+Z_{3,-1}) \biggr] \nonumber \\ && \quad +\left(\frac{\alpha_{s}}{\pi}\right)^3 \biggl[ \frac{1}{\varepsilon^3}(\Pi_{4,-3} -2\beta_0\Pi_{3,-2} +\beta_0^2\Pi_{2,-1} +Z_{4,-3}) \nonumber \\ && \quad \hspace{16mm} +\frac{1}{\varepsilon^2}(\Pi_{4,-2}-2\beta_0\Pi_{3,-1} +\beta_0^2\Pi_{2,0} - \beta_1\Pi_{2,-1}/2+Z_{4,-2}) \nonumber \\ && \quad \hspace{16mm} +\frac{1}{\varepsilon}(\Pi_{4,-1}-2\beta_0\Pi_{3,0} +\beta_0^2\Pi_{2,1} -\beta_1\Pi_{2,0}/2+Z_{4,-1})\biggr]\hspace{2mm} \equiv 0. \label{eq:PiPiexpans} \end{eqnarray} The leading poles in $Z_{\Pi}$ are absent at each order of $\alpha_s$ ($Z_{2,-2}=Z_{3,-3}=Z_{4,-4}=0$) except the zeroth order. Taking into account eq.\ (\ref{eq:Pirelat1}), we obtain the other set of relations between the perturbative coefficients of $\Pi$, $Z$ and QCD $\beta$-function \begin{displaymath} 3Z_{3,-2}+\beta_0Z_{2,-1}=0, \end{displaymath} \begin{equation} 2Z_{4,-3}+\beta_0Z_{3,-2}=0, \label{eq:Pirelat2} \end{equation} \begin{displaymath} 4Z_{4,-2}+2\beta_0Z_{3,-1}+\beta_1Z_{2,-1}=0. \end{displaymath} \vspace{3mm} \begin{displaymath} \Pi_{1,-1}=-Z_{1,-1}, \end{displaymath} \begin{displaymath} \Pi_{2,-1}=-Z_{2,-1}, \end{displaymath} \begin{displaymath} \Pi_{3,-2}=-Z_{3,-2}-\beta_0Z_{2,-1}, \end{displaymath} \begin{equation} \Pi_{3,-1}=-Z_{3,-1}+\beta_0\Pi_{2,0}, \label{eq:Pirelat3} \end{equation} \begin{displaymath} \Pi_{4,-1}=-Z_{4,-1}+2\beta_0\Pi_{3,0}+\beta_1\Pi_{2,0}/2-\beta_0^2\Pi_{2,1}, \end{displaymath} \begin{displaymath} \Pi_{4,-2}=-Z_{4,-2}-2\beta_0Z_{3,-1}-\beta_1Z_{2,-1}/2+\beta_0^2\Pi_{2,0}, \end{displaymath} \begin{displaymath} \Pi_{4,-3}=-Z_{4,-3}-2\beta_0Z_{3,-2}-\beta_0^2Z_{2,-1}. \end{displaymath} In section 6, the above relations will be used in the calculations of the four-loop total cross-section in electron-positron annihilation. \vspace{2cm} \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf Method for evaluation of renormalization constants} We now discuss the evaluation of renormalization constants within 't~Hooft's MS scheme ('t~Hooft, 1973), using Vladimirov's method (Vladimirov, 1978) and the so-called infrared rearrangement procedure (Vladimirov, 1980; Chetyrkin and Tkachov, 1982). To calculate the renormalizaton constant $Z_{\Gamma}$ for the one-particle-irreducible Green's function $\Gamma$, it is convenient to use the following representation (Vladimirov, 1978) \begin{equation} Z_{\Gamma}=1-{\cal K}R'\Gamma. \label{eq:Kdef} \end{equation} The operator $\cal K$ picks out all singular terms from the Laurent series in $\varepsilon$ \begin{displaymath} {\cal K}\sum_{i}c_{i}\varepsilon^i=\sum_{i<0}c_{i}\varepsilon^i. \end{displaymath} $R'$ is defined by the recursive relation \begin{equation} R'G=G-\sum_{G_i} {\cal K}R'G_{1}...{\cal K}R'G_{n}\times G_{/(G_{1}\cup...\cup G_{n})}, \label{eq:KR'} \end{equation} where the sum runs over all sets of one-particle-irreducible divergent subgraphs $G_i$ of the diagram $G$. $G_{/(G_{1}+...+G_{n})}$ is the diagram G with the subgraphs $G_{1},...,G_{n}$ shrunk to a point. In fact, $R'$ is the ordinary Bogolyubov-Parasyuk $R$-operation (Bogolyubov and Parasyuk, 1955a,b, 1956, 1957; for a textbook see Bogolyubov and Shirkov, 1980) without the last (overall) subtraction. Thus, $R'$ subtracts all ``internal'' divergences only and is connected to the ordinary $R$-operation in the following way \begin{displaymath} R=(1-{\cal K})R'. \end{displaymath} To calculate the renormalization constant $Z$ in eq.\ (\ref{eq:PiR}), one should write a diagram representation of $\Pi$ and apply ${\cal K}R'$ to the corresponding graphs ( eq.\ (\ref{eq:Kdef}) ) or, in other words, one should evaluate the counterterms for each graph. The benefit of using relation (\ref{eq:Kdef}) is based on the fact that the ${\cal K}R'$ for each diagram is a polynomial in dimensional parameters (Collins, 1974; Speer 1974). This fundamental property of the 't~Hooft's minimal subtraction prescription is the basic idea of the various versions of the infrared rearrangement technique (Vladimirov, 1980; Chetyrkin and Tkachov, 1982). As an example, we demonstrate the application of the ${\cal K}R'$ operation to the three-loop QCD diagram contributing to the $O(\alpha_{s}^2)$ total cross section for the process $e^{+}e^{-} \rightarrow \mbox{hadrons}$. \newpage \begin{displaymath} R'\biggl\{ \hspace{2cm} \biggr\} = \hspace{2cm} -2{\cal K}R'\biggl\{ \hspace{1cm} \biggr\}\hspace{1.5cm} -2{\cal K}R'\biggl\{ \hspace{1cm} \biggr\}\hspace{1.5cm} \end{displaymath} \begin{displaymath} +\biggl({\cal K}R'\biggl\{ \hspace{1cm} \biggr\}\biggr)^2 \end{displaymath} \vspace{3mm} \begin{displaymath} {\cal K}R'\biggl\{ \hspace{1cm} \biggr\}={\cal K}\biggl( \hspace{1cm} -{\cal K}R'\biggl\{ \hspace{1cm} \biggr\}\hspace{1.5cm} \biggr), \end{displaymath} \vspace{3mm} \begin{displaymath} {\cal K}R'\biggl\{ \hspace{1cm} \biggr\}={\cal K}\biggl\{ \hspace{1cm} \biggr\}. \end{displaymath} \vspace{3mm} The benefit of using the ${\cal K}R'$ operation besides its convenience in actual calculations is as follows. Using the fact that the result of ${\cal K}R'$ operation is a polynomial in masses and external momenta of the diagram, one can remove the dependence on the external momenta by differentiating (usually twice is sufficient) with respect to the external momentum and then setting the external momentum to zero. However, in this case infrared divergences appear. In order to prevent this, one can introduce a new fictitious external momentum as an infrared regulator flowing along some of the lines of the diagram (Chetyrkin and Tkachov, 1982). Alternatively, one can introduce a fictitious mass in one of the lines of the diagram as an infrared regulator (Vladimirov, 1980). An appropriate choice of the fictitious momentum can drastically simplify the topology of the given diagram. Both versions of the so called infrared rearrangement procedure simplify the calculation and make it possible to evaluate counterterms to four- and five-loop diagrams. The main result of the application of the infrared rearrangement technique can be formulated as follows. The problem of calculating the counterterms of an arbitrary $l$-loop diagram with an arbitrary number of masses and external momenta within the MS prescription can be reduced to the problem of calculating some $l-1$ -loop massless integrals to $O(\varepsilon^0)$ with only one external momentum. In the later sections, the full calculational procedure will be demonstrated for a typical four-loop diagram contributing to the photon renormalization constant. \vspace{2cm} \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf Evaluation of Wilson coefficient functions in operator product expansion} \indent In this subsection we briefly discuss the problem of evaluation of higher twist operator contributions to the hadronic vacuum polarization function. Those contributions are relevant in the analysis of nonperturbative contributions in some processes (e.g., hadronic decay of the $\tau$-lepton). We use the Wilson operator product expansion technique (Wilson, 1969) - mathematical apparatus allowing a factorization of the short distance contributions, which are calculable perturbatively and large distance effects which can be parametrized with the vacuum condensates ( Shifman, Vainshtein and Zakharov, 1979; Novikov {\it et al.}, 1985). In the perturbative evaluation of Wilson coefficient functions, we rely on the so called method of projectors (Gorishny, Larin and Tkachov, 1983; Gorishny and Larin, 1987; see also Pivovarov and Tkachov 1988, 1993 and references therein). An actual calculation for the coefficient functions of the operators of $\mbox{dim}=4$ has been done in the work by Loladze, Surguladze and Tkachov (1984, 1985), and Surguladze and Tkachov (1989b, 1990). The present discussion is based mainly on those works. Below we demonstrate the above technique in the case of the coefficient functions of gluon and quark condensates. Consider the operator product expansion of the T-product of two quark currents in the deep Euclidean region, $-q^2=Q^2 \rightarrow \infty$ \begin{equation} {\cal T}(Q)=i\int d^4x e^{iqx} T J(x)J(0) = \sum_{i} C_{i}(Q)O_{i}(0), \label{eq:OPE} \end{equation} where $J$ are quark currents. $C_{i}(Q)$ are c-number coefficient functions containing all dependence on $Q$. $O_{i}$ are local operators forming in general a complete basis. If the currents $J$ are gauge invariant then, after averaging over the vacuum, only gauge invariant operators contribute to the r.h.s. of eq.\ (\ref{eq:OPE}). However, the renormalization procedure mixes gauge invariant operators with non-invariant ones and one has to consider the complete basis of operators of the given dimension. The following set of operators of the dimension 4 \begin{displaymath} O_{1} = (G_{\mu\nu}^{a})^2, \hspace{3mm} O_{2}^{f}=m_{f}\overline{q}_{f}q_{f}, \hspace{3mm} O_{3}^f = \overline{q}_{f}(i\hat{\partial}-m_f+gT^{a}\hat{A^a})q_f, \end{displaymath} \begin{equation} O_{4}=(\partial_{\mu}\overline{c}^a)(\partial_{\mu}c^a) +(\partial_{\mu}\delta^{ab}+gf^{abc}A_{\mu}^{c})A_{\nu}^{b}G_{\mu\nu}^{a} -g\sum_{f}\overline{q}_{f}T^{a}\hat{A}^{a}q_f, \label{eq:OPEbasis} \end{equation} \begin{displaymath} O_{5} =\partial_{\mu}\overline{c}^a((\partial_{\mu}\delta^{ab} +gf^{abc}A_{\mu}^{c})c^{b} \end{displaymath} is closed under renormalization together with the ``operator'' $\sim m^4$ (Spiridonov, 1984; Loladze, Surguladze and Tkachov, 1984, 1985). Our aim is to calculate coefficient functions of gauge invariant operators $O_1$ and $O_{2}^{f}$. Note that $\sim m^4$ operators can be ignored because of the special structure of the renormalization matrix for the basis (\ref{eq:OPEbasis}). The Feynman rules for the operators (\ref{eq:OPEbasis}) are (Surguladze and Tkachov, 1990) \newpage \begin{displaymath} \lefteqn{O_1} \hspace{6cm} 4\delta^{ab}(p^2g^{\mu\nu}-p^{\mu}p^{\nu}) \end{displaymath} \vspace{2mm} \begin{displaymath} \lefteqn{O_{2}^{f}} \hspace{8cm} \delta_{ff'}m_f \end{displaymath} \vspace{2mm} \begin{displaymath} O_{3}^{f} \hspace{65mm} \delta_{ff'}(\hat{p}-m_f) \end{displaymath} \vspace{2mm} \begin{equation} O_4 \hspace{55mm} 2\delta^{ab}(p^2g^{\mu\nu}-p^{\mu}p^{\nu}) \label{eq:OPEFrules} \end{equation} \vspace{2mm} \begin{displaymath} O_4 \hspace{73mm} \delta^{ab}p^2g^{\mu\nu} \end{displaymath} \vspace{2mm} \begin{displaymath} O_5 \hspace{73mm} \delta^{ab}p^2g^{\mu\nu} \end{displaymath} \vspace{2mm} \begin{displaymath} O_5 \hspace{76mm} if^{abc}p^{\mu} \end{displaymath} \vspace{11mm} The operators of the basis (\ref{eq:OPEbasis}) are renormalized as follows \begin{equation} O_{i} = (Z_O)_{ij}O_{j}^{\mbox{\scriptsize B}}, \label{eq:OPEbasisren} \end{equation} where the superscript B marks the same operators as in (\ref{eq:OPEbasis}) but built from the ``bare'' fields, masses and couplings. The structure of the renormalization matrix $Z_O$ has been studied by Spiridonov (1984). In the MS type schemes $Z_O$ has the following form (Surguladze and Tkachov, 1990) \vspace{3cm} \begin{equation} \mbox{} \label{eq:OPEZ} \end{equation} \vspace{23mm} \noindent where only the matrix elements $A$ and $B$ are relevant. \begin{displaymath} A = \biggl(1-\frac{\beta(\alpha_s)}{\varepsilon}\biggr)^{-1}, \end{displaymath} \begin{equation} B = \frac{4\gamma_{m}(\alpha_s)}{\varepsilon} \biggl(1-\frac{\beta(\alpha_s)}{\varepsilon}\biggr)^{-1}. \label{eq:OPEAB} \end{equation} Inserting eq.\ (\ref{eq:OPEbasisren}) into the expansion (\ref{eq:OPE}) we get \begin{equation} {\cal T}(Q) = \sum_{i,j}C_{i}(Q)O_{i}^{\mbox{\scriptsize B}}(Z_O)_{ij}. \label{eq:PiRexpans} \end{equation} Following the method of projectors (Gorishny, Larin and Tkachov, 1983), we define the projectors $\pi_{i}$ satisfying the orthogonality condition and vanishing on higher spin operators \begin{equation} \pi_{i}[O_{j}^{\mbox{\scriptsize B}}]=\delta_{ij}, \label{eq:piproject} \end{equation} \begin{displaymath} \pi_{i}[\mbox{higher spin operators}]=0. \end{displaymath} Projectors $\pi_{i}$ applied on the l.h.s.\ of eq.\ (\ref{eq:PiRexpans}) separate in the r.h.s.\ the coefficient functions we are interested in \begin{equation} \pi_{j}[{\cal T}(Q)] = \sum_{i}C_{i}(Q)(Z_O)_{ij}. \label{eq:pionPi} \end{equation} We find the coefficient functions \begin{equation} C_{i}(Q)=\sum_{j}\pi_{j}[{\cal T}(Q)](Z_O^{-1})_{ji}. \label{eq:CF} \end{equation} Our aim is to find the coefficient functions of gauge invariant operators $O_1=(G_{\mu\nu})^2$ and $O_{2}^{f}=m_{f}\overline{q}_{f}q_f$. So, we need to construct the corresponding projectors $\pi_1$ and $\pi_{2}^{f}$. Let us represent $\pi_i$ as a linear combinations of some ``elementary'' projectors ${\cal P}_j$ defined in the following way. \begin{displaymath} {\cal P}_{1}[O] = \frac{1}{N_{A}}\frac{\partial^2}{\partial p^2} \delta^{ab}g^{\mu\nu} \biggl\{ \hspace{45mm} \biggr\}_{p=m_{f}=0} \end{displaymath} \vspace{1mm} \begin{displaymath} {\cal P}_{2}^{f}[O] = \frac{1}{4N_{F}}\frac{\partial}{\partial m_f} Tr\biggl\{ \hspace{45mm} \biggr\}_{p=m_{f}=0} \end{displaymath} \vspace{2mm} \begin{equation} \hspace{-2mm} {\cal P}_{3}^{f}[O] = \frac{1}{4N_{F}}\frac{\partial}{\partial p^{\sigma}} Tr \gamma^{\sigma}\biggl\{ \hspace{45mm} \biggr\}_{p=m_{f}=0} \label{eq:ElProj} \end{equation} \vspace{2mm} \begin{displaymath} \hspace{4mm} {\cal P}_{4}[O] = \frac{1}{N_{A}}\frac{\partial^2}{\partial p^2} \delta^{ab} \biggl\{ \hspace{45mm} \biggr\}_{p=m_{f}=0} \end{displaymath} \vspace{2mm} \begin{displaymath} \hspace{2mm} {\cal P}_{5}[O] = \frac{if^{abc}}{gN_{A}C_{A}}\frac{\partial} {\partial p^{\mu}} \biggl\{ \hspace{45mm} \biggr\}_{p=m_{f}=0} \end{displaymath} \vspace{1mm} \noindent where the parentheses contain the one-particle-irreducible Green function with one operator insertion. In the case of ${\cal P}_{2}^{f}$ and ${\cal P}_{3}^{f}$ the traces are calculated over Lorentz spinor and color indices. Acting by the projectors ${\cal P}_{j}$ on the operators (\ref{eq:OPEbasis}) we obtain \begin{displaymath} {\cal P}_{1}[O_{1}]=8D(D-1), \hspace{6mm} {\cal P}_{1}[O_{4}]=4D(D-1), \hspace{6mm} {\cal P}_{2}^{f}[O_{2}^{f'}]=\delta_{ff'}, \end{displaymath} \begin{equation} {\cal P}_{2}^{f}[O_{3}^{f'}]=-\delta_{ff'}, \hspace{6mm} {\cal P}_{3}^{f}[O_{3}^{f'}]=D\delta_{ff'}, \hspace{6mm} {\cal P}_{4}[O_{4}]={\cal P}_{4}[O_{5}]=2D, \hspace{6mm} {\cal P}_{5}[O_{5}]=D. \label{eq:P(o)} \end{equation} The results which are not shown in the above list are identicaly zero. {}From the definition (\ref{eq:piproject}) and eq.\ (\ref{eq:P(o)}) we obtain the explicit form for the projectors $\pi_{1}$ and $\pi_{2}^{f}$ \begin{displaymath} \pi_{1} = \frac{1}{8D(D-1)}[{\cal P}_{1}-2(D-1){\cal P}_{4}+4(D-1) {\cal P}_{5}], \end{displaymath} \begin{equation} \pi_{2}^{f}={\cal P}_{2}^{f}+\frac{1}{D}{\cal P}_{3}^{f}. \label{eq:piexplicit} \end{equation} Combining eqs.\ (\ref{eq:OPEZ}) and (\ref{eq:OPEAB}) with eq.\ (\ref{eq:CF}) we get our final expressions for the coefficient functions $C_{1}(Q)$ and $C_{2}^{f}(Q)$ (Surguladze and Tkachov, 1989,1990) \begin{displaymath} C_{1}(Q)=\pi_{1}[{\cal T}(Q)]\biggl(1-\frac{\beta(\alpha_s)}{\varepsilon}\biggr), \end{displaymath} \begin{equation} C_{2}^{f}(Q) =\pi_{2}^{f}[{\cal T}(Q)]-\pi_{1}[{\cal T}(Q)]\frac{4\gamma_{m}(\alpha_s)} {\varepsilon}. \label{eq:CFexplicit} \end{equation} The above expressions have a closed form and are valid at any order of perturbation theory. We note that ${\cal T}$ must be constructed with unrenormalized couplings and fields before one applies the projectors $\pi_{i}$. The general theory of Euclidean asymptotic expansions of Feynman integrals and the methods applicable to high order perturbative calculations have been developed in the works of Tkachov (1983b, 1983c, 1991, 1993), Chetyrkin and Tkachov (1982), and Chetyrkin (1991) (see also Smirnov 1990, 1991 and references therein). The technique developed in these works allows one to derive operator product expansions in the MS-scheme for any Feynman integral. For more general discussion and further details we refer to the above works and also to the original calculations (Surguladze and Tkachov, 1989, 1990). In section 5 we present a short description of the calculation of the coefficient functions of gluon and quark condensates up to $O(\alpha_{s}^2)$. \vspace{2cm} \renewcommand{\thesection}{\Roman{section}} \section{\bf \ $\Gamma(H\rightarrow hadrons)$ to $O(\alpha^2_s)$} \renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0} \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf The decay rate in terms of running parameters} \indent In this subsection, using the above methods we calculate the $O(\alpha^{2}_{s})$ corrections to the total hadronic decay width of the Standard Model Higgs boson in the massless quark limit (Gorishny, Kataev, Larin and Surguladze, 1990, 1991b; Surguladze, 1994b). \vspace{6cm} \begin{center} FIG.\ 1.\hspace{2mm} The process $H \rightarrow$ hadrons \end{center} \vspace{5mm} The standard SU(2)$\times$U(1) Lagrangian density of fermion-Higgs interaction is \begin{equation} L = -g_{Y}\overline{q}_fq_fH =-(\sqrt{2}G_F)^{1/2}m_f\overline{q}_fq_fH =-(\sqrt{2}G_F)^{1/2}j_fH. \label{eq:Lagr} \end{equation} The decay width of a scalar Higgs boson to the quark-antiquark pair is determined by the imaginary part of the two-point correlation function \begin{equation} \Pi(Q^2=-s,m_f)=i\int e^{iqx}<Tj_{f}(x)j_{f}(0)>_0d^4x \label{eq:pifunctions} \end{equation} of the quark scalar currents $j_f=m_f\overline{q}_fq_f$ in the following way \begin{equation} \Gamma_{H\rightarrow q_f\overline{q}_f} =\frac{\sqrt{2}G_F}{M_H} \mbox{Im}\Pi(s+i0,m_f)\biggr|_{s=M_H^2}. \label{eq:ImPi} \end{equation} $M_H$ is the Higgs mass. The total decay width will be the sum over all participating (depending on $M_H$) quark flavors \begin{equation} \Gamma(H \rightarrow \mbox{\small hadrons}) = \sum_{f=u,d,s,...} \Gamma_{H\rightarrow q_f\overline{q}_f}. \label{eq:total} \end{equation} We follow the work by Gorishny, Kataev, Larin and Surguladze (1990) and in analogy to the vector channel introduce the Adler function (Adler, 1974) \begin{equation} D(Q^2,m_f) = Q^2\frac{d}{dQ^2}\frac{\Pi(Q^2,m_f)}{Q^2}. \label{eq:Dfunctscalar} \end{equation} The derivative avoids the additive renormalization of $\Pi$. In fact, it is possible to proceed without the introduction of $D$-function and deal directly with the correlation function $\Pi$ (Surguladze, 1994b). Indeed, we are interested in $\mbox{Im}\Pi(s+i0,m_f)$. Since the overall MS renormalization constant has no terms like $(\log\mu^2/Q^2)^n/\varepsilon^k$, its imaginary part vanishes identically. The abscence of the pole logarithms in renormalization constants is a general feature of MS type schemes. The $D$-function obeys the homogeneous renormalization group equation \begin{equation} \left(\mu^2\frac{\partial}{\partial\mu^2} +\beta(\alpha_s)\alpha_s\frac{\partial}{\partial\alpha_s} -\gamma_m(\alpha_s)\frac{\partial} {\partial \log m_f}\right)D(\mu^2/Q^2,m_f,\alpha_s)=0. \label{eq:RGEDscalar} \end{equation} The QCD $\beta$-function and the mass anomalous dimension $\gamma_m$ are known up to the three loop approximation and have been given in the previous section. The plan for evaluation of $\Gamma_{H\rightarrow q_f\overline{q}_f}$ is as follows. First, we write the diagram representation for $\Pi(Q^2,m_f)$ according to the standard Feynman rules up to the desired loop-level. Second, we evaluate the Feynman diagrams using the dimensional regularization and renormalize the coupling and quark masses within the MS renormalization prescription. Finally, to get the decay rate, we analytically continue the result for the $D$-function obtained from eq.\ (\ref{eq:Dfunctscalar}) from Euclidean to Minkowski space. Following the above plan, we now demonstrate the calculation of $\Gamma_{H\rightarrow q_f\overline{q}_f}$ up to the 3-loop level. First of all, note that the correlation function $\Pi$ and the related $D$-function depend on quark masses. The algorithms for evaluation of the 3-loop Feynman diagrams constructed with the propagators of massive particles has not yet been developed. However, in the deep Euclidean region ($Q^2\rightarrow \infty$) it is possible to simplify the calculation using the expansion in terms of the small parameter $m_f^2/Q^2$ \begin{equation} \frac{1}{m_f^2Q^2}\Pi(Q^2,m_f) =\Pi(Q^2)+O\biggl(\frac{m_f^2}{Q^2}\biggr). \label{eq:Piexpa} \end{equation} Such an expansion is legitimate since we consider a Higgs boson much heavier than the typical hadronic mass scale. In this section we calculate the first term in the above expansion and the related decay rate. This is equivalent to the assumption that all five quarks are massless and the top quark decouples ($m_t \rightarrow \infty$). The diagrammatic representation for $\Pi$ in somewhat symbolic form looks like \vspace{7mm} \begin{displaymath} \Pi(Q^2) \sim \hspace{2cm} + \frac{\alpha_s}{\pi}\biggl[\hspace{23mm}+ \hspace{3mm} 2\hspace{23mm} \biggr] \end{displaymath} \vspace{3mm} \begin{equation} +\biggl(\frac{\alpha_s}{\pi}\biggr)^2\biggl[\hspace{23mm} +\cdots +(\mbox{total of 16 three-loop diagrams})\biggr] +O(\alpha^{3}_{s}). \label{eq:pidiagram} \end{equation} \vspace{7mm} \noindent Next, we evaluate one-, two- and three-loop massless Feynman diagrams. By simple power counting, it is easy to find that in general the above diagrams are UV divergent. The unrenormalized contribution from a typical three-loop diagram in the $\overline{\mbox{MS}}$ renormalization scheme (Bardeen, Buras, Duke and Muta, 1978) reads \vspace{3cm} \begin{displaymath} \sim \frac{1}{(4\pi)^2} \left(\frac{\alpha_{s}^{\mbox{\tiny B}}}{4\pi}\right)^{2}N_{F} \frac{C_{F}C_{A}}{2} (m_{f}^{\mbox{\tiny B}})^{2}Q^2 \left(\frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{Q^2}\right)^{3\varepsilon} \biggl[ \frac{16}{\varepsilon^3} +\frac{400}{3\varepsilon^2} +\frac{2344}{3\varepsilon} -\frac{160}{\varepsilon}\zeta(3) +\frac{11800}{3} \end{displaymath} \begin{displaymath} \hspace{9cm} -1312\zeta(3)-240\zeta(4)+320\zeta(5) \biggr], \end{displaymath} where $m_{f}^{\mbox{\tiny B}}$ is the $f$-flavor quark mass originating from the quark mass dependence of the Yukawa coupling. $\zeta(3)$, $\zeta(4)$ and $\zeta(5)$ are ordinary Riemann $\zeta$-functions. The number 2 in front of the diagram stands for the symmetry factor. The algorithms for the evaluation of propagator type one-, two- and three-loop massless Feynman diagrams have been given by Tkachov (1981, 1983a) and Chetyrkin and Tkachov (1981). For the description of the algorithms see also Gorishny, Larin, Surguladze and Tkachov, 1989. The results given in this section were reobtained with the help of the program {\small HEPL}oops (Surguladze, 1992) and the previous results (Gorishny, Kataev, Larin and Surguladze, 1990, 1991) were independently confirmed (Surguladze, 1994b). As one can see, each three-loop diagram in general may contain a pole with power $\leq 3$. In the vector channel, after summing the results for all diagrams with an appropriate symmetry and SU(N) group factor, the leading pole cancels. This is the consequence of the conservation of electromagnetic currents. In the scalar channel, the leading poles remain in $\Pi$. This is related to the quark mass dependence of the coupling. Evaluating the unrenormalized correlation function (\ref{eq:pifunctions}) and using the definition (\ref{eq:Dfunctscalar}), we obtain the unrenormalized $D$-function in the massless limit. \begin{eqnarray} \lefteqn{\hspace{-9mm} D\left( \frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2},\alpha_s\right) = \frac{1}{(4\pi)^2}N_{F}(m_{f}^{\mbox{\tiny B}})^2 \biggl\{ \left(\frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{Q^2}\right)^{\varepsilon} (2+4\varepsilon+8\varepsilon^2) } \nonumber\\ && \quad \hspace{7mm} +\left(\frac{\alpha_{s}^{\mbox{\tiny B}}}{4\pi}\right) \left(\frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{Q^2}\right)^{2\varepsilon} C_F\biggl[\frac{12}{\varepsilon}+58+\varepsilon(227-48\zeta(3))\biggr] \nonumber\\ && \quad \hspace{6mm} +\left(\frac{\alpha_{s}^{\mbox{\tiny B}}}{4\pi}\right)^2 \left(\frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{Q^2}\right)^{3\varepsilon} C_F\biggl[ \hspace{5mm} C_{F}\biggl( \frac{36}{\varepsilon^2}+\frac{279}{\varepsilon} +\frac{3139}{2}-360\zeta(3) \biggr) \nonumber\\ && \quad \hspace{48mm} +C_{A}\biggl( \frac{22}{\varepsilon^2}+\frac{201}{\varepsilon} +\frac{2511}{2}-300\zeta(3) \biggr) \nonumber\\ && \quad \hspace{48mm} -TN \biggl( \frac{8}{\varepsilon^2}+\frac{68}{\varepsilon} +414-96\zeta(3) \biggr) \biggr]+O(\alpha_{s}^{3}) \biggr\}. \label{eq:Dbare} \end{eqnarray} The above expression requires the renormalization of the strong coupling ( eq.\ (\ref{eq:AsR}) ) and the multiplicative renormalization ( eq.\ (\ref{eq:mrenorm}) ) originating from the quark mass dependence of the Yukawa coupling. Expanding the factors $(\mu_{\overline{\mbox{\tiny MS}}}^{2}/Q^2)^{l\varepsilon}$ in terms of $\varepsilon$ and performing the renormalizations of the coupling and the quark mass, we get a finite analytical expression for the $D$-function in the $\overline{\mbox{MS}}$ scheme \begin{eqnarray} \lefteqn{\hspace{-5mm} D\left( \frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2},\alpha_s\right)= \frac{N_F}{8\pi^2}m_{f}^{2}\biggl\{1 +\biggl(\frac{\alpha_s}{4\pi}\biggr)C_F \biggl[17 +6\log \biggl(\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr) \biggr]} \nonumber\\ && \quad +\biggl(\frac{\alpha_s}{4\pi}\biggr)^2C_F \biggl[ C_F\biggl(\frac{691}{4}-36\zeta(3)\biggr) +C_A\biggl(\frac{893}{4}-62\zeta(3)\biggr) -TN(65-16\zeta(3)) \nonumber\\ && \quad \hspace{4cm} +\log\biggl(\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr) \biggl( 105C_F+\frac{284}{3}C_A-\frac{88}{3}TN \biggr) \nonumber\\ && \quad \hspace{4cm} +\log^2\biggl(\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr) ( 18C_F+11C_A-4TN ) \biggr] \biggr\}. \label{eq:Danalyt0} \end{eqnarray} For standard QCD with the color SU$_{\mbox{\scriptsize c}}$(3) symmetry group, the analytical result for the $D$-function reads (Surguladze, 1989d) \begin{eqnarray} \lefteqn{D\left( \frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2},\alpha_s\right)=} \nonumber\\ && \frac{3}{8\pi^2}m_{f}^{2}\biggl\{1 +\biggl(\frac{\alpha_s}{\pi}\biggr)\biggl[\frac{17}{3} +2\log\biggl(\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr) \biggr] \nonumber\\ && \quad \hspace{13mm} +\biggl(\frac{\alpha_s}{\pi}\biggr)^2 \biggl[ \frac{10801}{144}-\frac{39}{2}\zeta(3) -\biggl(\frac{65}{24}-\frac{2}{3}\zeta(3)\biggr)N \\ && \quad \hspace{24mm} +\log\biggl(\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr) \biggl(\frac{106}{3}-\frac{11}{9}N \biggr) +\log^2\biggl(\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr) \biggl(\frac{19}{4}-\frac{1}{6}N\biggr) \biggr] \biggr\}. \nonumber \label{eq:Danalyt} \end{eqnarray} This completes the evaluation of the correlation function of the two scalar quark currents in the massless limit at the three-loop approximation. There is one crucial test of this calculation based on the renormalization group constraints. The solution of the renormalization group equation (\ref{eq:RGEDscalar}) can be conveniently rewritten as follows \begin{equation} D\biggl(\frac{\mu^2}{Q^2},m_{f}(\mu),\alpha_s(\mu)\biggr)= \frac{3}{8\pi^2} m_f^2(\mu)\sum_{0\leq j \leq i} \biggl(\frac{\alpha_s(\mu)}{\pi}\biggr)^i a_{ij} \log^j\frac{\mu^2}{Q^2}. \label{eq:gamma0gen} \end{equation} Applying the differential operator $\mu^2 d/d\mu^2$ to both sides of eq.\ (\ref{eq:gamma0gen}), taking into account the renormalization group invariance of the $D$-function and eqs.\ (\ref{eq:beta}) and (\ref{eq:gamma}), we obtain to $O(\alpha_s)$ \begin{equation} a_{11}=2\gamma_0a_{00}, \label{eq:relations1} \end{equation} to $O(\alpha_s^2)$ \begin{displaymath} a_{21}=2\gamma_1a_{00}+(\beta_0+2\gamma_0)a_{10}, \end{displaymath} \begin{equation} a_{22}=(\beta_0+2\gamma_0)\frac{a_{11}}{2}=(\beta_0+2\gamma_0)\gamma_0a_{00}, \label{eq:relations2} \end{equation} and to $O(\alpha_s^3)$ \begin{displaymath} a_{31}=2(\beta_0+\gamma_0)a_{20}+(\beta_1+2\gamma_1)a_{10}+2\gamma_2a_{00}, \end{displaymath} \begin{displaymath} a_{32}=(\beta_0+\gamma_0)a_{21}+(\beta_1+2\gamma_1)\frac{a_{11}}{2} =(\beta_0+\gamma_0)[2\gamma_1a_{00}+(\beta_0+2\gamma_0)a_{10}] +(\beta_1+2\gamma_1)\gamma_0a_{00}, \end{displaymath} \begin{equation} a_{33}=\frac{2}{3}(\beta_0+\gamma_0)a_{22}=\frac{2}{3}\gamma_0 (\beta_0+\gamma_0)(\beta_0+2\gamma_0)a_{00}. \label{eq:relations3} \end{equation} The relations (\ref{eq:relations1}) and (\ref{eq:relations2}) provide a powerful check of our calculation, while the relations (\ref{eq:relations3}) allow one to evaluate the $\log$ terms to $O(\alpha_s^3)$, without explicit calculations of the corresponding four-loop diagrams. With those relations, the information available at present, namely the QCD $\beta$-function, mass anomalous dimension and the two-point correlation function up to the three-loop level is fully exploited. In fact, similar relations can be derived for the correlation function $\Pi$. However, the renormalization group equation for $\Pi$ is not a homogeneous one and the anomalous dimension function up to the corresponding order of $\alpha_s$ is necessary. We evaluate the decay rate of the neutral Higgs boson into a quark antiquark pair by analytical continuation of $D\left(\mu^{2}/Q^2,m_{f}(\mu),\alpha_s(\mu)\right)$ from Euclidean to Minkowski space. The total decay rate can be obtained by summing up over all participating quark flavors. \begin{eqnarray} \lefteqn{\hspace{-7mm}\Gamma(H\rightarrow \mbox{\small hadrons}) =\frac{3\sqrt{2}G_FM_H}{8\pi} \sum_{f=u,d,s,...} m_f^2\biggl\{ 1+\frac{\alpha_s}{\pi}\biggl(\frac{17}{3} +2\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{M_H^2}\biggr)}\nonumber\\ && \quad +\biggl(\frac{\alpha_s}{\pi}\biggr)^2 \biggl[\frac{10801}{144}-\frac{19}{2}\zeta(2)-\frac{39}{2}\zeta(3) +\frac{106}{3}\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{M_H^2} +\frac{19}{4}\log^2 \frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{M_H^2}\nonumber\\ && \quad \hspace{1cm} -N\biggl(\frac{65}{24}-\frac{1}{3}\zeta(2)-\frac{2}{3}\zeta(3) +\frac{11}{9}\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{M_H^2} +\frac{1}{6}\log^2 \frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{M_H^2}\biggr)\biggr] \biggr\}. \label{eq:GHtot} \end{eqnarray} The Riemann function $\zeta(2)=\pi^2/6$ arose from the analytical continuation of the $\log^2\mu_{\overline{\mbox{\tiny MS}}}^2/Q^2$ term and $\zeta(3)=1.202056903$. The procedure of analytical continuation and the appearance of invariant additional contributions have been discussed in several earlier works (Krasnikov and Pivovarov, 1982; Pennington and Ross, 1982; Radyushkin, 1982; Pivovarov, 1992a). Note that in some cases those additional corrections are large and affect the result significantly. This is especially true for the total cross section in the process $e^{+}e^{-} \rightarrow \mbox{hadrons}$. To minimize such corrections it was proposed, for instance, to redefine the expansion parameter (Pennington and Ross, 1982; Radyushkin, 1982). \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf The decay rate in terms of pole quark mass} For the heavy flavor decay mode of the Higgs, it is relevant to parametrize the decay rate in terms of quark pole mass (see, e.g., Kniehl, 1994a). Let us rewrite the result for $\Gamma_{H\rightarrow q_f\overline{q}_f}$ in terms of pole quark mass, assuming that heavy quark is not exactly on-shell. This subsection is based mainly on recent findings (Surguladze, 1994a,b). Solving the renormalization group equation for the quark mass - eq.\ (\ref{eq:RGfunctions}), we obtain the following scaling law for the running quark mass: \begin{equation} \frac{m_f(\mu_1)}{m_f(\mu_2)}=\frac{\phi(\alpha_s(\mu_1))} {\phi(\alpha_s(\mu_2))}, \label{eq:mrun} \end{equation} where \begin{eqnarray} \lefteqn{\phi(\alpha_s(\mu))=\biggl(2\beta_0\frac{\alpha_s(\mu)}{\pi}\biggr) ^{\frac{\gamma_0}{\beta_0}} \biggl\{1 +\biggl(\frac{\gamma_1}{\beta_0} -\frac{\beta_1\gamma_0}{\beta_0^2}\biggr)\frac{\alpha_s(\mu)}{\pi}} \nonumber\\ && \quad \hspace{-3mm} +\frac{1}{2}\biggl[\biggl(\frac{\gamma_1}{\beta_0} -\frac{\beta_1\gamma_0}{\beta_0^2}\biggr)^2 +\frac{\gamma_2}{\beta_0} -\frac{\beta_1\gamma_1}{\beta_0^2}-\frac{\beta_2\gamma_0}{\beta_0^2} +\frac{\beta_1^2\gamma_0}{\beta_0^3} \biggr]\biggl(\frac{\alpha_s(\mu)}{\pi}\biggr)^2\biggr\}. \label{eq:f} \end{eqnarray} In the above equation all appropriate quantities are evaluated for $N$ active quark flavors. $N$ can be determined according to the scale of $M_H$. At present one usually considers $N=5$. For the running coupling we obtain the following evolution equation to $O(\alpha_s^3)$ (Surguladze, 1994b) \begin{eqnarray} \lefteqn{\hspace{-1cm}\frac{\alpha_s^{(n)}(\mu_1)}{\pi} =\frac{\alpha_s^{(N)}(\mu_2)}{\pi}} \nonumber\\ && +\biggl(\frac{\alpha_s^{(N)}(\mu_2)}{\pi}\biggr)^2 \biggl(\beta_0^{(N)}\log\frac{\mu_2^2}{\mu_1^2} +\frac{1}{6}\sum_{l}\log\frac{m_l^2}{\mu_1^2} \biggr) \nonumber\\ && +\biggl(\frac{\alpha_s^{(N)}(\mu_2)}{\pi}\biggr)^3 \biggl[\beta_1^{(N)}\log\frac{\mu_2^2}{\mu_1^2} +\frac{19}{24}\sum_{l}\log\frac{m_l^2}{\mu_1^2} \nonumber\\ && \quad \hspace{23mm} +\biggl(\beta_0^{(N)}\log\frac{\mu_2^2}{\mu_1^2} +\frac{1}{6}\sum_{l}\log\frac{m_l^2}{\mu_1^2} \biggr)^2 \nonumber\\ && \quad \hspace{23mm} -\frac{25}{72}(N-n)\biggr], \label{eq:Astransform} \end{eqnarray} where the superscript $N$ ($n$) indicates that the corresponding quantity is evaluated for $N$ ($n$) numbers of participating quark flavors. Conventionally (see, e.g., Marciano, 1984) $N$ ($n$) is specified to be the number of quark flavors with mass $\leq \mu_2$ ($\leq \mu_1$). However, eq.\ (\ref{eq:Astransform}) is relevant for any $n\leq N$ and arbitrary $\mu_1$ and $\mu_2$, regardless of the conventional specification of the number of quark flavors. The $\log m_l/\mu_1$ terms are due to the ``quark treshold'' crossing effects and the constant coefficients $1/6=\beta_0^{(k-1)}-\beta_0^{(k)}$, $19/24=\beta_1^{(k-1)}-\beta_1^{(k)}$ represent the contributions of the quark loop in the $\beta$-function. The sum runs over $N-n$ quark flavors (e.g., $l=b$ if $n=4$ and $N=5$). Note that $m_l$ is the pole mass of the quark with flavor $l$. For the on-shell definition of the quark masses eq.\ (\ref{eq:Astransform}) changes - the constant $-25/72$ should be substituted by $+7/72$. The above equation is derived based on eq.\ (\ref{eq:Asparametr}), the QCD matching conditions for $\alpha_s$ at ``quark thresholds'' (Bernreuter and Wetzel, 1982; Marciano, 1984; Barnett, Haber and Soper, 1988; Rodrigo and Santamaria, 1993) and the one-loop relation between on-shell and pole quark masses. Eq.(\ref{eq:Astransform}) is consistent with the QCD matching relation at $m_f(m_f)$ (Bernreuter and Wetzel, 1982) \begin{equation} \alpha_s^{(N_f-1)}(m_f(m_f))=\alpha_s^{(N_f)}(m_f(m_f)) +(\alpha_s^{(N_f)}(m_f(m_f)))^3(C_A/9-17C_F/96)/\pi^2 \label{eq:2lmatch} \end{equation} Here and below $N_f$ is the number of quark flavors $u,d,...,f$. Note that the nonlogarithmic constant at $O(\alpha_s^3)$ in eq.\ (\ref{eq:Astransform}) will not contribute in further analysis. Next, using the scaling properties of the MS running mass and eq.\ (\ref{eq:Astransform}), one obtains the following matching condition \begin{eqnarray} \lefteqn{\hspace{-9mm}m_f^{(N-1)}(\mu)=m_f^{(N)}(\mu)\biggl\{1+ \biggl(\frac{\alpha_s^{(N)}(\mu)}{\pi}\biggr)^2 \biggl[\delta(m_f,m_{f'})-\frac{5}{36}\log\frac{\mu^2}{m_f^2} -\frac{1}{12}\log^2\frac{\mu^2}{m_f^2}} \nonumber\\ && \quad \hspace{63mm} +\frac{1}{6}\log\frac{\mu^2}{m_f^2}\log\frac{\mu^2}{m_{f'}^2} -\frac{2}{9}\log\frac{m_{f'}^2}{m_f^2}\biggr]\biggr\} \label{eq:massmatch}, \end{eqnarray} where the constant terms are: $1/12=\gamma_0(\beta_0^{(k-1)}-\beta_0^{(k)})/2$, $5/36=\gamma_1^{(k-1)}-\gamma_1^{(k)}$ and $2/9=C_F(\beta_0^{(k-1)}-\beta_0^{(k)})$. In general, the $\delta(m_f,m_{f'})$ is the finite contribution of the single virtual heavier quark with mass $m_{f'}$, entering when one increases the number of flavors from $N-1$ to $N$ (one can also consider the particular case $m_{f'}=m_f$). {}From the two-loop on-shell quark mass renormalization one has (Broadhurst, Gray and Schilcher, 1991) \begin{equation} \delta(m_f,m_{f'})=-\zeta(2)/3-71/144 +(4/3)\Delta(m_{f'}/m_f), \label{eq:deltaM} \end{equation} where \begin{equation} \Delta(r)=\frac{1}{4}\biggl[\log^2r+\zeta(2)-\biggl(\log r +\frac{3}{2}\biggr)r^2 -(1+r)(1+r^3)L_{+}(r)-(1-r)(1-r^3)L_{-}(r)\biggr], \label{eq:delta} \end{equation} \begin{displaymath} L_{\pm}(r) = \int_{0}^{1/r}dx\frac{\log x}{x \pm 1}. \end{displaymath} $L_{\pm}(r)$ can be evaluated for different quark mass ratios $r$ numerically. We relate the $\overline{\mbox{MS}}$ quark mass $m_f(m_f)$ to the pole mass $m_f$ using the $O(\alpha_s^2)$ on-shell results of Broadhurst, Gray and Schilcher (1991) \begin{equation} m_f^{(N_f)}(m_f)=m_f\biggl[1-\frac{4}{3}\frac{\alpha_s^{(N_f)}(m_f)}{\pi} +\biggl(\frac{16}{9}-K_{f}\biggr) \biggl(\frac{\alpha_s^{(N_f)}(m_f)}{\pi}\biggr)^2\biggr], \label{eq:mtopole} \end{equation} where \begin{equation} K_f= \frac{3817}{288}+\frac{2}{3}(2+\log2)\zeta(2)-\frac{1}{6}\zeta(3) -\frac{N_f}{3}\biggl(\zeta(2)+\frac{71}{48}\biggr) +\frac{4}{3}\sum_{m_l \leq m_f} \Delta\biggl(\frac{m_l}{m_f}\biggr). \label{eq:K} \end{equation} The first four terms in $K_f$ represent the QCD contribution with $N_f$ massless quarks, while the sum is the correction due to the $N_f$ nonvanishing quark masses. Combining eqs.\ (\ref{eq:mrun}), (\ref{eq:f}) and eqs. (\ref{eq:Astransform})-(\ref{eq:mtopole}), one obtains the relation between the $\overline{\mbox{MS}}$ quark mass $m_f(M_H)$ renormalized at $M_H$ and evaluated for the $N$-flavor theory and the pole quark mass $m_f$ (Surguladze, 1994b) \begin{eqnarray} \lefteqn{m_f^{(N)}(M_H)=m_f\biggl\{1 -\frac{\alpha_s^{(N)}(M_H)}{\pi} \biggl(\frac{4}{3}+\gamma_0\log\frac{M_H^2}{m_f^2}\biggr)} \nonumber\\ && \quad \hspace{-7mm} -\biggl(\frac{\alpha_s^{(N)}(M_H)}{\pi}\biggr)^2\biggl[K_f +\sum_{m_f<m_{f'}<M_H}\delta(m_f,m_{f'}) -\frac{16}{9} +\biggl(\gamma_1^{(N)}-\frac{4}{3}\gamma_0 +\frac{4}{3}\beta_0^{(N)}\biggr)\log\frac{M_H^2}{m_f^2} \nonumber\\ && \quad \hspace{19mm} +\frac{\gamma_0}{2}(\beta_0^{(N)}-\gamma_0)\log^2\frac{M_H^2}{m_f^2} \biggr]\biggr\}. \label{eq:mMHtopole} \end{eqnarray} Note that $N$ is specified according to the size of $M_H$ and has no correlation with the quark mass $m_f$. Thus, for instance, one can apply eq.\ (\ref{eq:mMHtopole}) to the charm mass $m_c^{(5)}(M_H)$ evaluated for five-flavor theory. Substituting eqs. (\ref{eq:mMHtopole}), (\ref{eq:K}) and appropriate $\beta$-function and mass anomalous dimension coefficients (see section 2) into eq.\ (\ref{eq:GHtot}), one obtains the decay rate in terms of the pole quark masses \begin{eqnarray} \lefteqn{\Gamma(H\rightarrow \mbox{\small hadrons}) =\frac{3\sqrt{2}G_FM_H}{8\pi}\sum_{f=u,d,s,...}m_f^2\biggl\{1 +\frac{\alpha_s^{(N)}(M_H)}{\pi} \biggl(3-2\log\frac{M_H^2}{m_f^2}\biggr)} \nonumber\\ && \quad +\biggl(\frac{\alpha_s^{(N)}(M_H)}{\pi}\biggr)^2 \biggl[\frac{697}{18} -\biggl(\frac{73}{6}+\frac{4}{3}\log 2\biggr)\zeta(2) -\frac{115}{6}\zeta(3) -N\biggl(\frac{31}{18}-\zeta(2)-\frac{2}{3}\zeta(3)\biggr) \nonumber\\ && \quad -\biggl(\frac{87}{4}-\frac{13}{18}N\biggr)\log\frac{M_H^2}{m_f^2} -\biggl(\frac{3}{4}-\frac{1}{6}N\biggr)\log^2\frac{M_H^2}{m_f^2} -\frac{8}{3}\sum_{m_l<M_H}\Delta\biggl(\frac{m_l}{m_f}\biggr) \biggr]\biggr\}. \label{eq:Polemassresf} \end{eqnarray} Recall, that at the beginning we have neglected terms which are suppressed by powers $m_f^2/M_H^2$. Such corrections to the decay rate, in general, may not be entirely negligible and have to be taken into account in precise numerical analyses. Presently those corrections due to the nonvanishing quark masses have also been calculated. For the explicit results, we refer to the original works (Surguladze, 1994a,b; Kniehl, 1995a; Chetyrkin and Kwiatkowski, 1995). In the next section we give the results for the quark mass corrections to the correlation functions $\Pi$. The full analytical result for the decay rate of $H \rightarrow q_f\overline{q}_f$ in terms of pole quark masses, including the leading order (two-loop) QCD corrections has been obtained independently by several groups: Braaten and Leveille (1980), Inami and Kubota (1981), and Dreess and Hikasa (1990). In the work by Sakai (1980) the two-loop result has been obtained in the zero quark mass limit. \begin{equation} \Gamma_{H\rightarrow q_f\overline{q}_f} =\frac{3\sqrt{2}G_FM_H}{8\pi}m_f^2 \biggl(1-\frac{4m_f^2}{M_H^2}\biggr)^{\frac{3}{2}} \biggl[1+\frac{\alpha_s(M_H)}{\pi} \delta^{(1)}(\frac{m_f^2}{M_H^2}) +O(\alpha_s^2)\biggr], \label{eq:2loop} \end{equation} where \begin{displaymath} \delta^{(1)}=\frac{4}{3}\biggl[\frac{a(\eta)}{\eta} +\frac{3+34\eta^2-13\eta^4}{16\eta^3}\log\omega +\frac{21\eta^2-3}{8\eta^2}\biggr], \end{displaymath} \begin{displaymath} a(\eta)=(1+\eta^2)\biggl[4Li_2(\omega^{-1})+2Li_2(-\omega^{-1}) -\log\omega\log\frac{8\eta^2}{(1+\eta)^3}\biggr] -\eta\log\frac{64\eta^4}{(1-\eta^2)^3}, \end{displaymath} \begin{displaymath} \omega=\frac{1+\eta}{1-\eta}, \hspace{5mm} \eta=\biggl(1-\frac{4m_f^2}{M_H^2}\biggr)^{\frac{1}{2}} \end{displaymath} and the Spence function is defined as usual \begin{displaymath} Li_2(x) = -\int_{0}^{x}dx\frac{\log(1-x)}{x} =\sum_{n=1}^{\infty}\frac{x^n}{n^2}. \end{displaymath} The expansion of the r.h.s of eq.\ (\ref{eq:2loop}) in a power series in terms of small $m_f^2/M_H^2$ has the following form \begin{eqnarray} \lefteqn{\hspace{-9mm}\Gamma_{H\rightarrow q_f\overline{q}_f} =\frac{3\sqrt{2}G_FM_H}{8\pi}m_f^2 \biggl\{\biggl(1-6\frac{m_f^2}{M_H^2}+...\biggr)} \nonumber\\ && \quad \hspace{-5mm} +\frac{\alpha_s(M_H)}{\pi} \biggl[3-2\log\frac{M_H^2}{m_f^2} -\frac{m_f^2}{M_H^2}\biggl(8-24\log\frac{M_H^2}{m_f^2}\biggr) +...\biggr]+O(\alpha_s^2)\biggr\}, \label{eq:2loopexpan} \end{eqnarray} where the periods cover higher order terms $\sim (m_f/M_H)^{2k}$, $k=2,3...$ One can see that the leading terms agree with the result (\ref{eq:Polemassresf}). Numerically the $\overline{\mbox{MS}}$ high order QCD corrections for the considered process are large and reduce the decay rates by about 40\%. \vspace{6mm} \renewcommand{\thesection}{\Roman{section}} \section{\bf Quark mass corrections to the correlation functions} \renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0} In the previous section we neglected all quark masses in the corresponding Feynman diagrams in comparison with the momentum scale of the problem. In other words, we have calculated the leading term in the expansion in terms of small $m_f^2/s$ (for the Higgs boson decay, $s=M_H^2$) in the limit of infinitely heavy top quark, $m_t\rightarrow \infty$. However, in the real world quarks are massive and the leading term in the above expansion may not always give a satisfactory approximation. On the other hand, starting at $O(\alpha_s^2)$, virtual heavy quark can also appear in certain topological type of Feynman diagrams (Fig.\ 2 and Fig.\ 3) regardless of the momentum scale of the problem. \vspace{32mm} \begin{center} FIG.\ 2. \ $O(\alpha_s^2)$ Feynman diagrams responsible for the virtual heavy quark contribution \end{center} \vspace{32mm} \begin{center} FIG.\ 3. \ $O(\alpha_s^2)$ Feynman diagrams responsible for the contribution due to the top-bottom mass splitting \end{center} According to the decoupling theorem (Appelquist and Carazzone, 1975), virtual quarks much heavier than the momentum scale of the problem decouple. However, for instance, in the process of Z boson decay the effect of the top quark may not be entirely negligible since $m_t$ is not much greater than $M_Z$. A similar role could be played by the charm quark in the hadronic decay of the tau-lepton. The evaluation of the virtual top quark contribution (Fig.\ 2) to the decay rate $Z\rightarrow \mbox{hadrons}$ and related quantities has been done in Kniehl (1990), Soper and Surguladze (1994), and Hoang, Jezabek, K\"{u}hn and Teubner (1994) without using large or small mass approximations. The correction turned out to be moderate and in good agreement with the results obtained with the help of the large mass expansion technique (Chetyrkin, 1993a). The contribution of the diagrams in Fig.\ 2, in the presence of a virtual heavy quark, to the two-point correlation function of the electromagnetic quark currents has been evaluated previously by Wetzel and Bernreuther (1981). Kniehl and K\"{u}hn (1989, 1990) have calculated the $O(\alpha_s^2)$ correction to the decay rate $Z\rightarrow \mbox{hadrons}$ due to the large mass splitting in the top-bottom doublet (Fig.\ 3). This correction turned out to be large and important. In this section we consider only the leading correction in the expansion in terms of small quark mass. For the calculations of virtual heavy quark contributions we refer the reader to the above mentioned original works (see also Kniehl, 1994b, 1995b). The discussion in this section is based on the works by Surguladze (1994a,b,c). Let us expand the full two-point correlation function, defined by eq.\ (\ref{eq:pifunction}) in the vector channel and by eq.\ (\ref{eq:pifunctions}) in the scalar and pseudoscalar channels, in powers of $m_f^2/Q^2$ in the ``deep'' Euclidean region \begin{equation} \biggl(\frac{1}{m_f^2Q^2}\biggr)^d\Pi(Q^2,m_f,m_{\mbox{\tiny V}}) =\Pi_1(Q^2)+\frac{m_f^2}{Q^2}\Pi_{m_f^2}(Q^2) +\sum_{{\mbox{\tiny V}} =u,d,s,c,b}\frac{m_{\mbox{\tiny V}}^2}{Q^2}\Pi_{m_{\mbox{\tiny V}}^2}(Q^2) +..., \label{eq:Piexpan} \end{equation} where $d=0$ in the vector channel and $d=1$ in the scalar and pseudoscalar channels. The last term in the above expansion is due to the Feynman diagrams containing a virtual fermionic loop. Note however that in the vector channel the contribution from the diagrams in Fig.\ 3 vanishes according to Furry's theorem (Furry, 1937). In order to evaluate the coefficient functions in the r.h.s of eq.\ (\ref{eq:Piexpan}), it is sufficient to write the diagrammatic representation for $\Pi(Q^2,m_f^{\mbox{\tiny B}},m_{\mbox{\tiny V}}^{\mbox{\tiny B}})$ up to the desired level of perturbation theory and apply the appropriate projector. To $O(\alpha_s^2)$ one has \begin{eqnarray} \lefteqn{\hspace{-9mm}\Pi_{m_{f}^{2n} m_{\mbox{\tiny V}}^{2k}}(Q^2,\alpha_s) =} \nonumber\\ && \frac{1}{(2n)!(2k)!} \biggl(\frac{d}{dm_f^{\mbox{\tiny B}}}\biggr)^{2n} \biggl(\frac{d}{dm_{\mbox{\tiny V}}^{\mbox{\tiny B}}}\biggr)^{2k} \biggl\{\frac{\Pi(Q^2,m_f^{\mbox{\tiny B}} ,m_{\mbox{\tiny V}}^{\mbox{\tiny B}},\alpha_s^{\mbox{\tiny B}})} {(m_f^{\mbox{\tiny B}})^{2d} Q^{2(d-n-k)}}\biggr\} _{\stackrel{m_f^{\mbox{\tiny B}}=m_{\mbox{\tiny V}}^{\mbox{\tiny B}}=0} {\alpha_s^{\mbox{\tiny B}}\rightarrow Z_{\alpha}\alpha_s}} (Z_m^2)^{(1+d)}, \label{eq:Proj} \end{eqnarray} where $n,k=0,1$, $n+k\leq 1$, and superscript ``B'' denotes the bare quantities. The mass renormalization constant $Z_m=m_f^{\mbox{\tiny B}}/m_f$ can be obtained from eq.\ (\ref{eq:mrenorm}). The Feynman diagrams contributing to the $\Pi_{m_{f}^{2n}m_{\mbox{\tiny V}}^{2k}}$ are the same as for the calculation of $\Pi_1$ (see eq.\ (\ref{eq:pidiagram})) but with massive fermion propagators. The calculations of all one-, two- and three-loop diagrams have been done using the program {\small HEPL}oops (Surguladze, 1992). The obtained expressions for $\Pi_i$ at each order of $\alpha_s$ are polynomials with respect to $1/\varepsilon$ and $\log\mu_{\overline{\mbox{\tiny MS}}}^2/Q^2$. The poles can be removed by an additive renormalization. We note that there are no terms like $(1/\varepsilon^n)(\log\mu_{\overline{\mbox{\tiny MS}}}^2/Q^2)^k$. They appear only at higher orders $\sim m_f^2 m_f^4/Q^4$ and represent infrared mass logarithms. The corresponding prescription similar to the Bogolyubov ultraviolet $R$-operation has been worked out in the work by Chetyrkin, Gorishny and Tkachov (1982), Tkachov (1983b,c), and Gorishny, Larin and Tkachov (1983). (see also Tkachov, 1991, 1993 and references therein). The infrared mass singularities have been studied earlier by Marciano (1975). In the present paper we consider only the terms $\sim m_f^2/Q^2$ which are sufficient for most of the phenomenologically interesting applications. In the vector channel we obtain the following $\overline{\mbox{MS}}$ analytical result (Gorishny, Kataev and Larin, 1986; Surguladze, 1994c) \begin{eqnarray} \lefteqn{\hspace{-9mm}\Pi_{m_{f}^2}\left( \frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2},\alpha_s\right)= \frac{N_F}{(4\pi)^2}\biggl\{-8 -\biggl(\frac{\alpha_s}{\pi}\biggr)C_F \biggl(16 +12\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr)} \nonumber\\ && \quad -\biggl(\frac{\alpha_s}{\pi}\biggr)^2 \biggl[ C_F^2\biggl(\frac{1667}{24} -\frac{5}{3}\zeta(3)-\frac{70}{3}\zeta(5) +\frac{51}{2}\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +9\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} \biggr) \nonumber\\ && \quad \hspace{1cm} +C_FC_A\biggl(\frac{1447}{24} +\frac{16}{3}\zeta(3)-\frac{85}{3}\zeta(5) +\frac{185}{6}\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{11}{2}\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} \biggr) \nonumber\\ && \quad \hspace{1cm} -C_FTN\biggl(\frac{95}{6} +\frac{26}{3}\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +2\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} \biggr) \biggr]\biggr\}, \label{eq:PimanalytV1} \end{eqnarray} \begin{equation} \hspace{-66mm}\Pi_{m_{\mbox{\tiny V}}^2} \left(\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2},\alpha_s\right) =\frac{N_F}{(4\pi)^2}\biggl(\frac{\alpha_s}{\pi}\biggr)^2C_FT \biggl[\frac{64}{3}-16\zeta(3)\biggr]. \label{eq:PimanalytV2} \end{equation} The contribution to the physical process, in particular to the decay rate of $Z\rightarrow \mbox{hadrons}$ can be obtained simply by taking the imaginary part in the r.h.s. of eqs.\ (\ref{eq:PimanalytV1}) and (\ref{eq:PimanalytV2}) at $Q^2=-s+i0$. We note, that the $\Pi_{m_{f}^2}$ and $\Pi_{m_{\mbox{\tiny V}}^2}$ turned out to be finite. No overall subtraction is necessary. Moreover, one can see that the imaginary part or the contribution to the decay rate vanishes at the parton level. This can be checked by the calculation of the parton contribution in the vector channel with explicit dependence on quark mass. Indeed, calculating the trivial fermionic loop we obtain \begin{equation} \Pi_{\mbox{\scriptsize parton}} (\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{-q^2},\frac{m_f^2}{-q^2}) =\frac{N_F}{(4\pi^2)} \biggl[\frac{4}{3}\frac{1}{\varepsilon} -8\int_{0}^{1}x(1-x)\log\frac{m_f^2-x(1-x)q^2} {\mu_{\overline{\mbox{\tiny MS}}}^2}dx \biggr]. \label{eq:1lPi} \end{equation} Taking the discontinuity under the integral and then evaluating the trivial integral with the $\Theta$ function, we obtain \begin{equation} \frac{1}{2\pi i}\mbox{disc}\Pi_{\mbox{\scriptsize parton}} (\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{-q^2},\frac{m_f^2}{-q^2}) =\frac{N_F}{(4\pi^2)}\biggl(1+\frac{2m_f^2}{q^2}\biggr) \sqrt{1-\frac{4m_f^2}{q^2}} =\frac{N_F}{(4\pi^2)}O\biggl(\frac{m_f^4}{q^4}\biggr). \label{eq:1lPidisc} \end{equation} The $\sim m_f^2/Q^2$ contribution to the Adler $D$-function can be obtained from eqs.\ (\ref{eq:PimanalytV1}) and (\ref{eq:PimanalytV2}) by differentiating with respect to $Q^2$. There is some confusion concerning the above results in the literature. Initially, the corrections $\sim m_f^2/Q^2$ in the vector channel have been calculated by Gorishny, Kataev and Larin (1986). Later, in the similar calculations (Surguladze, 1989a), a slightly different result was obtained, which was confirmed in further publications (see, e.g., Kataev, 1990, 1991). However, in the recent works (Chetyrkin and Kwiatkowski, 1993; Surguladze, 1994c), the initial result of Gorishny, Kataev and Larin (1986) has been confirmed. Unfortunately, in the analysis of the mass corrections to the $Z$ decay rates (Chetyrkin and K\"{u}hn, 1990) the incorrect result was used. Fortunately, the main conclusions of Chetyrkin and K\"{u}hn (1990) are not affected. Summarizing, we note that the results (\ref{eq:PimanalytV1}) and (\ref{eq:PimanalytV2}) (Gorishny, Kataev and Larin, 1986; Chetyrkin and Kwiatkowski, 1993; Surguladze, 1994c) seem now to be reliable. In the scalar channel the result for the standard SU$_{\mbox{\scriptsize c}}$(3) gauge group reads (Surguladze, 1994b) \begin{eqnarray} \lefteqn{\hspace{-1mm}\Pi_{m_f^2}\biggl( \frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{Q^2},\alpha_s\biggr)= -\frac{1}{4\pi^2}\biggl\{12+9 \log\frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{Q^2}} \nonumber\\ && \quad +\biggl(\frac{\alpha_s}{\pi}\biggr)\biggl(94-36\zeta(3) +60\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +18\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr) \nonumber\\ && \quad +\biggl(\frac{\alpha_s}{\pi}\biggr)^2 \biggl[\frac{17245}{16} -\frac{1690}{3}\zeta(3)-3\zeta(4)+\frac{385}{3}\zeta(5) \nonumber\\ && \quad \hspace{9mm} +\biggl(\frac{7149}{8}-249\zeta(3)\biggr) \log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{1113}{4}\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{81}{2}\log^3\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} \nonumber\\ && \quad \hspace{9mm} -N\biggl(\frac{817}{24}-6\zeta(3) +\biggl(\frac{313}{12}-6\zeta(3)\biggr) \log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{15}{2}\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\log^3\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr)\biggr] \nonumber\\ && \quad \hspace{77mm} +\mbox{\small ``simple poles''} \biggr\}, \label{eq:PimanalytS1} \end{eqnarray} \begin{equation} \hspace{-42mm}\Pi_{m_{\mbox{\tiny V}}^2} \left(\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2},\alpha_s\right) =\frac{1}{4\pi^2}\biggl(\frac{\alpha_s}{\pi}\biggr)^2 \biggl[\frac{8}{3}+6\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\mbox{\small ``simple pole''}\biggr], \label{eq:PimanalytS2} \end{equation} where under the ``simple pole'' we mean number$/\varepsilon^k$ with no dependence on $\log\mu^2/Q^2$. The ``simple poles'' have no imaginary part and consequently will not contribute to the observable quantities at the given order of $\alpha_s$. Note that the $\Pi_{m_{\mbox{\tiny V}}^2}$ in eq. (\ref{eq:PimanalytS2}) does not include the contribution from the triangle anomaly type graphs pictured in Fig.\ 3. Those graphs make the following additional contribution to $\Pi$ in eq.\ (\ref{eq:Piexpan}) (Surguladze, 1994b) \begin{equation} +\sum_{f'=u,d,s,c,b}\frac{m_{f'}^2}{Q^2} \times \frac{1}{4\pi^2}\biggl(\frac{\alpha_s}{\pi}\biggr)^2 \biggl[\frac{118}{3}-20\zeta(3)-10\zeta(5) +12\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\mbox{\small ``simple pole''}\biggr]. \label{eq:PimanalytS3} \end{equation} The above results are relevant for the decay rate of the standard model Higgs boson into a quark antiquark pair, calculated in the previous section in the massless quark limit. Corrections $\sim m_f^2/M_H^2$ can be obtained from eqs.\ (\ref{eq:PimanalytS1}), (\ref{eq:PimanalytS2}) and (\ref{eq:PimanalytS3}) (Surguladze, 1994b). In the pseudoscalar channel we define the quark currents as $j_f=m_f\overline{q}_fi\gamma_5 q_f$. We also define the $\gamma_5$ matrix in $D$-dimensional space-time as an object with the following properties \begin{equation} \{\gamma_5,\gamma_\mu\}=0, \hspace{5mm} \gamma_5\gamma_5=1. \label{eq:gamma5def} \end{equation} The above definition causes no problems in dimensional regularization when there are two $\gamma_5$ matrices in a closed fermionic loop. We obtain (Surguladze, 1994a) \begin{eqnarray} \lefteqn{\hspace{-4mm}\Pi_{m_{f}^2}\left( \frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2},\alpha_s\right)= -\frac{1}{4\pi^2}\biggl\{3 \log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\biggl(\frac{\alpha_s}{\pi}\biggr)\biggl(6-12\zeta(3) +4\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +6\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr)} \nonumber\\ && \quad +\biggl(\frac{\alpha_s}{\pi}\biggr)^2 \biggl[-\frac{6713}{144} -116\zeta(3)-\zeta(4)+\frac{235}{3}\zeta(5) \nonumber\\ && \quad \hspace{1cm} +\biggl(\frac{1429}{24}-83\zeta(3)\biggr) \log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{155}{4}\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{27}{2}\log^3\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} \nonumber\\ && \quad \hspace{5mm} -N\biggl(-\frac{31}{72}-\frac{2}{3}\zeta(3) +\biggl(\frac{9}{4}-2\zeta(3)\biggr) \log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{7}{6}\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{1}{3}\log^3\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr)\biggr] \nonumber\\ && \quad \hspace{77mm} +\mbox{\small ``simple poles''} \biggr\}, \label{eq:PimanalytPS1} \end{eqnarray} \begin{equation} \hspace{-43mm}\Pi_{m_{\mbox{\tiny V}}^2} \left(\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2},\alpha_s\right) =\frac{1}{4\pi^2}\biggl(\frac{\alpha_s}{\pi}\biggr)^2 \biggl[\frac{8}{3}+6\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\mbox{\small ``simple pole''}\biggr]. \label{eq:PimanalytPS2} \end{equation} The result for the pseudoscalar channel is relevant, for instance, for the decay rates of the minimal supersymmetric version of the Higgs particle into a quark antiquark pair (see Surguladze, 1994a). Finally, we present the results of calculation of the $\sim m_f^2/Q^2$ corrections to the correlation function in the axial channel (Soper and Surguladze, 1994; Surguladze, 1994c). We use the following definition of the correlation function \begin{equation} i\int d^4x e^{iqx}<Tj_{\mu}^f(x)j_{\nu}^f(0)>_0= g_{\mu\nu}Q^2\Pi(Q,m_f)-Q_{\mu}Q_{\nu}\Pi'(Q,m_f), \label{eq:AxialPi} \end{equation} where $j_{\mu}^f=\overline{q}_f\gamma_{\mu}\gamma_5q_f$. Note that in the axial channel the correlation function is not transverse in contrast to the vector channel. However, for the decay rate of the $Z$-boson only the $\sim g_{\mu\nu}$ part in eq.\ (\ref{eq:AxialPi}) is relevant. The expansions of $\Pi$ and $\Pi'$ in terms of small $m_f^2/Q^2$ has the same form as in the vector channel ( eq.\ (\ref{eq:Piexpan}) ). The coefficient functions in this expansion can be calculated according to eq.\ (\ref{eq:Proj}) in the vector channel. In the calculations of one-, two- and three-loop Feynman diagrams the program {\small HEPL}oops (Surguladze, 1992) was used. The final results for the SU$_{\mbox{\scriptsize c}}$(3) gauge group read (Soper and Surguladze, 1994; Surguladze, 1994c) \newpage \begin{eqnarray} \lefteqn{\hspace{-2mm}\Pi_{m_{f}^2}\left( \frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2},\alpha_s\right)= \frac{1}{4\pi^2}\biggl\{6+6 \log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} } \nonumber\\ && \quad +\biggl(\frac{\alpha_s}{\pi}\biggr)\biggl(\frac{107}{2}-24\zeta(3) +22\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +6\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr) \nonumber\\ && \quad +\biggl(\frac{\alpha_s}{\pi}\biggr)^2 \biggl[\frac{3241}{6} -387\zeta(3)-\frac{3}{2}\zeta(4)+165\zeta(5) \nonumber\\ && \quad \hspace{1cm} +\biggl(\frac{8221}{24}-117\zeta(3)\biggr) \log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{155}{2}\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{19}{2}\log^3\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} \nonumber\\ && \quad \hspace{5mm} -N\biggl(\frac{857}{36}-\frac{32}{3}\zeta(3) +\biggl(\frac{151}{12}-4\zeta(3)\biggr) \log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{8}{3}\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} +\frac{1}{3}\log^3\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr)\biggr] \nonumber\\ && \quad \hspace{77mm} +\mbox{\small ``simple poles''} \biggr\} \label{eq:PimanalytAT1} \end{eqnarray} \begin{equation} \hspace{-79mm}\Pi_{m_{\mbox{\tiny V}}^2} \left(\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2},\alpha_s\right) =\frac{1}{4\pi^2}\biggl(\frac{\alpha_s}{\pi}\biggr)^2 \biggl[\frac{32}{3}-8\zeta(3)\biggr] \label{eq:PimanalytAT2} \end{equation} \begin{eqnarray} \lefteqn{\hspace{-22mm}\Pi'_{m_{f}^2}\left( \frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2},\alpha_s\right)= \frac{1}{4\pi^2}\biggl\{-6 +\biggl(\frac{\alpha_s}{\pi}\biggr)\biggl(-12 -12\log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr)} \nonumber\\ && \quad +\biggl(\frac{\alpha_s}{\pi}\biggr)^2 \biggl[-\frac{4681}{24} -34\zeta(3)+115\zeta(5) \nonumber\\ && \quad \hspace{1cm} -\frac{215}{2} \log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} -\frac{57}{2}\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} \nonumber\\ && \quad \hspace{9mm} -N\biggl(-\frac{55}{12}-\frac{8}{3}\zeta(3) -\frac{11}{3} \log\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2} -\log^2\frac{\mu_{\overline{\mbox{\tiny MS}}}^{2}}{Q^2}\biggr)\biggr] \nonumber\\ && \quad \hspace{67mm} +\mbox{\small ``simple poles''} \biggr\} \label{eq:PimanalytAL1} \end{eqnarray} \begin{equation} \Pi'_{m_{\mbox{\tiny V}}^2}=\Pi_{m_{\mbox{\tiny V}}^2} \label{eq:PimanalytAL2} \end{equation} The results given in this section can be tested using the renormalization group. Namely, the relations similar to eqs.\ (\ref{eq:relations1}), (\ref{eq:relations2}) and (\ref{eq:relations3}) can be obtained here (Surguladze, 1994a,b,c). In fact, in the vector channel, one can obtain the $O(\alpha_s^3)$ logarithmic terms without actual calculation of the corresponding four-loop diagrams. On the other hand, the leading logarithmic terms in $\Pi$-function form the corresponding contribution to the decay rates of, for instance, the Z-boson (Chetyrkin and K\"{u}hn, 1990; Chetyrkin, K\"{u}hn and Kwiatkowski, 1992; Surguladze, 1994c). In the axial channel the situation is more complicated. Here, because the renormalization group equation similar to eq.\ (\ref{eq:RGEDscalar}) is no longer a homogeneous one, the renormalization group approach is restricted to $O(\alpha_s^2)$. \renewcommand{\thesection}{\Roman{section}} \section{\bf Two-loop coefficient functions of $\mbox{dim}=4$ power corrections} \renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0} In this section we outline the calculations of the two-loop coefficient functions of $\mbox{dim}=4$ power corrections. We consider the contributions which appear in the short distance expansion of the correlation function of two flavor-diagonal vector, scalar and pseudoscalar currents constructed from light quark fields. The methods and corresponding references are given in the earlier sections. The corrections for the vector channel have been evaluated in Loladze, Surguladze and Tkachov (1984, 1985) and Surguladze and Tkachov (1988). In the scalar and pseudoscalar channels, the calculation has been done in Surguladze and Tkachov (1990). The calculation for vector and axial vector channels has been done in Chetyrkin, Gorishny and Spiridonov (1985), where the previous results for the vector channel have been confirmed and the calculation was extended for flavor non-diagonal currents as well. The three-loop correction to the coefficient function of gluon condensate in the scalar channel has also been computed in Surguladze and Tkachov (1989b). For the calculation of dimension 8 terms in the operator product expansion see also Broadhurst and Generalis (1985). Here we follow the work by Surguladze and Tkachov (1990). Consider first the T-product of flavor diagonal vector currents of light quarks \begin{equation} {\cal T}_{\mu\nu}^{f'}(Q)=i\int d^{4}x e^{iqx} T J_{\mu}^{f'}(x)J_{\nu}^{f'}(0), \label{eq:Pid4V} \end{equation} where $J^{f'}_{\mu}=\overline{q}_{f'}\gamma_{\mu}q_{f'}$. Taking into account the current conservation and operator product expansion technique (Wilson, 1969) for large momentum transfer ($Q^2 \rightarrow \infty$) we write \begin{equation} {\cal T}_{\mu\nu}^{f'}(Q)=(g_{\mu\nu}Q^2-Q_{\mu}Q_{\nu}) \biggl\{C_{0}+\frac{1}{Q^4}\biggl[C_{G^2}(Q^2)(G_{\mu\nu}^{a})^2+ \sum_{f'}C_{\overline{q}q}^{f'}(Q^2)m_{f'}\overline{q}_{f'}q_{f'}\biggr] +\cdots \biggr\}, \label{eq:Pid4VOPE} \end{equation} where $C_{0}$ is the coefficient function of the unity operator including the terms $\sim m_f^2/Q^2$ discussed in the previous section. The period covers the operators of higher twists. For the scalar and pseudoscalar channels the transverse factor in the above equation is absent. To simplify the calculation, we contract over the Lorentz indices $\mu$ and $\nu$. Then the expressions for $C_{i}$ defined in eq. (\ref{eq:Pid4VOPE}) coincide with the ones in eq.\ (\ref{eq:CFexplicit}) if ${\cal T}(Q)$ is replaced by ${\cal T}_{\mu\mu}^{f'}(Q^2)/(D-1)Q^2$, where $D=4-2\varepsilon$. Let us rewrite eqs.\ (\ref{eq:CFexplicit}) in a somewhat symbolic diagrammatic representation to $O(\alpha_s^2)$. \vspace{3mm} \begin{displaymath} C_{G^2}=\pi_{1}\frac{\alpha_s}{\pi}\biggl\{2\hspace{16mm}+4\hspace{16mm} +\frac{\alpha_s}{\pi} \biggl[\hspace{17mm}+...+\hspace{17mm}+... +\hspace{17mm}+...\biggr]\biggr\} \end{displaymath} \vspace{3mm} \begin{displaymath} \hspace{-8mm}C_{\overline{q}q}^{f} =\pi_{2}^{f}\biggl\{2\hspace{23mm}+\frac{\alpha_s}{\pi} \biggl[2\hspace{23mm}+4\hspace{23mm}+2\hspace{23mm}+...\biggr] \end{displaymath} \vspace{2mm} \begin{equation} \hspace{93mm} +\biggl(\frac{\alpha_s}{\pi}\biggr)^2\biggl[\hspace{23mm}+...\biggr] \biggr\}. \label{eq:OPEdiagrammer} \end{equation} \vspace{3mm} The total number of two-loop graphs contributing to $C_{G^2}$ is 30 and to $C_{\overline{q}q}^{f}$ is 38. There is a simple rule for generating the appropriate graphs at $O(\alpha_{s}^n)$. One should take the graphs contributing to $O(\alpha_{s}^{(n+1)})$ in the unity operator and disconnect one fermion line in all possible ways for the coefficient function $C_{\overline{q}q}^{f}$. For the coefficient function $C_{G^2}$ it is necessary to write all the diagrams with one disconnected gluon line (relevant for the projector ${\cal P}_{1}$), all the diagrams with one disconnected ghost line (relevant for the projector ${\cal P}_{4}$) and all the diagrams with disconnected gluon-ghost-ghost vertex (relevant for the projector ${\cal P}_{5}$). To see this, recall eqs.\ (\ref{eq:ElProj}). Acting with the projectors ${\cal P}_{j}$ on the appropriate diagrams, the calculations are reduced to the evaluation of one- and two-loop propagator type massless Feynman integrals. In the original calculation (Loladze, Surguladze and Tkachov, 1984, 1985; Surguladze and Tkachov, 1988, 1990) all Feynman integrals have been evaluated analytically using the {\small REDUCE} (Hearn, 1973) program LOOPS (Surguladze and Tkachov, 1989a). The $\overline{\mbox{MS}}$ results for the projectors ${\cal P}_{j}$ in the vector channel read \begin{equation} {\cal P}_{1}[{\cal T}_{\mu\mu}^{f'}]=\frac{1}{Q^4}C_{F}\frac{N_{F}}{N_{A}} \frac{\alpha_{s}^{\mbox{\tiny B}}}{\pi}\biggl\{48-32\varepsilon +\frac{\alpha_{s}^{\mbox{\tiny B}}}{\pi}\biggl[C_{F}(-12) +C_{A}(\frac{18}{\varepsilon}-42+72\zeta(3)\biggr)\biggr] +O(\alpha_{s}^2)\biggr\} \label{eq:P1} \end{equation} \begin{equation} {\cal P}_{4}[{\cal T}_{\mu\mu}^{f'}]=\frac{1}{Q^4}C_{F}\frac{N_{F}}{N_{A}}C_{A} \biggl(\frac{\alpha_{s}^{\mbox{\tiny B}}}{\pi}\biggr)^2 \biggl(\frac{3}{\varepsilon}-9+12\zeta(3)\biggr) +O(\alpha_{s}^3), \label{eq:P4} \end{equation} \begin{equation} {\cal P}_{5}[{\cal T}_{\mu\mu}^{f'}]=0+O(\alpha_{s}^3), \label{eq:P5} \end{equation} \begin{equation} \biggl({\cal P}_{2}^{f\neq f'}+\frac{1}{D}{\cal P}_{3}^{f\neq f'} \biggr)[{\cal T}_{\mu\mu}^{f'}]= \frac{1}{Q^4}C_{F}T \biggl(\frac{\alpha_{s}^{\mbox{\tiny B}}}{\pi}\biggr)^2 \biggl(\frac{24}{\varepsilon}-60+96\zeta(3)\biggr) +O(\alpha_{s}^3), \label{eq:P231} \end{equation} \begin{eqnarray} \lefteqn{\hspace{-12mm} \biggl({\cal P}_{2}^{f=f'}+\frac{1}{D}{\cal P}_{3}^{f=f'}\biggr) [{\cal T}_{\mu\mu}^{f'}]=} \nonumber\\ && \quad \frac{1}{Q^4}\biggl\{6 +\frac{\alpha_{s}^{\mbox{\tiny B}}}{\pi}C_{F}\biggl(\frac{3}{2} +\frac{11}{4}\varepsilon\biggr) +\biggl(\frac{\alpha_{s}^{\mbox{\tiny B}}}{\pi}\biggr)^2C_{F} \biggl[C_{F}\frac{387}{16}+C_{A}\biggl(\frac{11}{8\varepsilon} +\frac{7}{16}\biggr) \nonumber\\ && \quad \hspace{7mm} +T\biggl(\frac{3}{4\varepsilon}-\frac{15}{4}+6\zeta(3)\biggr) -TN\biggl(\frac{1}{2\varepsilon}+\frac{7}{4}\biggr)\biggr] +O(\alpha_{s}^3) \biggr\}. \label{eq:P232} \end{eqnarray} The vanishing of ${\cal P}_{5}[{\cal T}_{\mu\mu}^{f'}]$ at the two-loop level is the consequence of gauge invariance, as was shown by Spiridonov (1987). Combining eqs.\ (\ref{eq:piexplicit}) and (\ref{eq:CFexplicit}) with the above results and renormalizing the bare coupling via eq. (\ref{eq:AsR}) we obtain $\overline{\mbox{MS}}$ $O(\alpha_{s}^2)$ analytical expressions for the coefficient functions in the vector channel (Surguladze and Tkachov, 1990) \begin{equation} C_{G^2}(Q^2)=\frac{1}{Q^4}C_{F}\frac{N_{F}}{N_{A}}\frac{1}{6} \frac{\alpha_{s}}{\pi}\biggl[1 +\frac{\alpha_{s}}{\pi}\biggl(\frac{C_{A}}{2}-\frac{C_{F}}{4}\biggr) +O(\alpha_{s}^2) \biggr], \label{eq:Cgganalytic} \end{equation} \begin{equation} C_{\overline{q}q}^{f=f'}(Q^2)=\frac{1}{Q^4} \biggl\{2+\frac{\alpha_{s}}{\pi}\frac{C_{F}}{2}\biggl[1 +\frac{\alpha_{s}}{\pi} \biggl(\frac{129}{8}C_{F}-\frac{25}{18}C_{A}-\frac{5}{9}TN +T(-3+4\zeta(3))\biggr)+O(\alpha_{s}^2)\biggr]\biggr\} \label{eq:Cqqanalytic1} \end{equation} \begin{equation} C_{\overline{q}q}^{f \neq f'}(Q^2)=\frac{1}{Q^4} \biggl(\frac{\alpha_{s}}{\pi}\biggr)^2C_{F}T \biggl(-\frac{3}{2}+2\zeta(3)\biggr)+O(\alpha_{s}^3). \label{eq:Cqqanalytic2} \end{equation} The above results are gauge invariant. This statement was checked by straightforward calculation in an arbitrary covariant gauge up to the term $\sim \varepsilon$ (Surguladze and Tkachov, 1990). The dependence on the gauge parameter canceled. Thus, it is simplest to do the calculation in the Feynman gauge. For simplicity, we have omitted the terms $\sim \log(\mu_{\overline{\mbox{\tiny MS}}}^2/Q^2)$, taking $\mu_{\overline{\mbox{\tiny MS}}}^2=Q^2$. The dependence on $\mu$ can be restored via the renormalization group (see below). Note that the coefficient function $C_{\overline{q}q}^{f \neq f'}$ is due to the diagrams pictured in Fig.\ 4 with disconnected fermion lines of the virtual loop (see also Fig.\ 2). \vspace{4cm} \begin{center} FIG.\ 4. \hspace{2mm} Two-loop diagrams forming $C_{\overline{q}q}^{f \neq f'}$ \end{center} Specifically for QCD with the SU$_{\mbox{\scriptsize c}}$(3) symmetry group we obtain \begin{equation} C_{G^2}(Q^2)=\frac{1}{Q^4}\frac{1}{12} \frac{\alpha_{s}}{\pi}\biggl(1+\frac{\alpha_{s}}{\pi}\frac{7}{6} +O(\alpha_{s}^2)\biggr), \label{eq:CgganalyticQCD} \end{equation} \begin{equation} C_{\overline{q}q}^{f=f'}(Q^2)=\frac{1}{Q^4} \biggl\{2+\frac{2}{3}\frac{\alpha_{s}}{\pi}\biggl[1 +\frac{\alpha_{s}}{\pi} \biggl(\frac{95}{6}+2\zeta(3)-\frac{5}{18}N\biggr) +O(\alpha_{s}^2)\biggr]\biggr\}, \label{eq:Cqqanalytic1QCD} \end{equation} \begin{equation} C_{\overline{q}q}^{f \neq f'}(Q^2)=\frac{1}{Q^4} \biggl(\frac{\alpha_{s}}{\pi}\biggr)^2 \biggl(-1+\frac{4}{3}\zeta(3)\biggr)+O(\alpha_{s}^3). \label{eq:Cqqanalytic2QCD} \end{equation} Note the very large $O(\alpha_{s}^2)$ coefficient in eq.\ (\ref{eq:Cqqanalytic1QCD}). However, this coefficient is renormalization scheme dependent and requires special analysis (see below). In the scalar and pseudoscalar channels the general expression for the coefficient functions eq.\ (\ref{eq:CF}) takes the form \begin{equation} C_{i}(Q)=Z_{m}^2\sum_{j}\pi_{j} \biggl[\frac{{\cal T}(Q)}{(m_{f}^{\mbox{\tiny B}})^2}\biggr](Z_{O}^{-1})_{ji}, \label{eq:CFscalar} \end{equation} where $Z_{m}=m_{f}^{\mbox{\tiny B}}/m_{f}$ is the quark mass renormalization constant (see eq.\ (\ref{eq:mrenorm}) ). The $\gamma^{5}$ matrix is defined within the dimensional regularization according to eq.\ (\ref{eq:gamma5def}). It is easy to see that in the calculations of $C_{G^2}$, two matrices $i\gamma^{5}$ can be anticommuted over the fermion propagators and ``annihilate'' each other so that the results in both channels coincide. The calculational procedure is exactly the same as it was for the vector channel, except for the need of mass renormalization. The results for the coefficient functions $C_{G^2}$ and $C_{\overline{q}q}^{f}$ in the $\overline{\mbox{MS}}$ -scheme are as follows (Surguladze and Tkachov, 1986, 1990). \noindent In the (pseudo)scalar channel \begin{equation} C_{G^2}(Q^2)=\frac{1}{Q^2}C_{F}\frac{N_{F}}{N_{A}}\frac{1}{4} \frac{\alpha_{s}}{\pi}\biggl[1 +\frac{\alpha_{s}}{\pi}\biggl(\frac{3}{2}C_{A} +\frac{3}{4}C_{F}\biggr) +O(\alpha_{s}^2) \biggr]. \label{eq:CgganalyticS} \end{equation} In the scalar channel \begin{eqnarray} \lefteqn{\hspace{-30mm}C_{\overline{q}q}^{f=f'}(Q^2)= \frac{1}{Q^2}\biggl\{3+\frac{\alpha_{s}}{\pi}\frac{39}{4}C_{F}\biggl(1 +\frac{\alpha_{s}}{\pi} \biggl[C_{F}\biggl(\frac{447}{208}-\frac{21}{13}\zeta(3)\biggr) +C_{A}\biggl(\frac{389}{144}+\frac{3}{26}\zeta(3)\biggr)} \nonumber\\ && \quad \hspace{35mm} -\frac{5}{39}T-\frac{25}{36}TN\biggr] +O(\alpha_{s}^2)\biggr)\biggr\}, \label{eq:Cqqanalytic1S} \end{eqnarray} \begin{equation} C_{\overline{q}q}^{f \neq f'}(Q^2)=\frac{1}{Q^2} \biggl(\frac{\alpha_{s}}{\pi}\biggr)^2C_{F}T \biggl(-\frac{5}{4}\biggr)+O(\alpha_{s}^3) \label{eq:Cqqanalytic2S} \end{equation} and in the pseudoscalar channel \begin{eqnarray} \lefteqn{\hspace{-30mm}C_{\overline{q}q}^{f=f'}(Q^2)= -\frac{1}{Q^2}\biggl\{1+\frac{\alpha_{s}}{\pi}\frac{17}{4}C_{F}\biggl(1 +\frac{\alpha_{s}}{\pi} \biggl[C_{F}\biggl(\frac{583}{272}-\frac{45}{17}\zeta(3)\biggr) +C_{A}\biggl(\frac{2443}{816}+\frac{27}{34}\zeta(3)\biggr)} \nonumber\\ && \quad \hspace{35mm} +\frac{5}{17}T-\frac{167}{204}TN\biggr] +O(\alpha_{s}^2)\biggr)\biggr\}. \label{eq:Cqqanalytic1PS} \end{eqnarray} The result for $C_{\overline{q}q}^{f \neq f'}$ coincides with the analogous one for the scalar channel. Let us turn to the renormalization group analysis of the above results. in this particular case it is possible to use the following trick (Surguladze and Tkachov, 1990). Note first that the vacuum average of the renormalized operators $G^2$ and $m\overline{q}{q}$ and their coefficient functions depend on the renormalization parameter $\mu$ and therefore are not convenient for further analysis. However, as was shown by Collins, Duncan and Joglekar (1977) (see also Nielsen, 1977; Tarrach, 1982; Narison and Tarrach, 1983), the vacuum average of the trace of the energy-momentum tensor \begin{equation} <\Theta_{\alpha\alpha}>_{0}=-\frac{\beta(\alpha_s)}{2\beta_{0}} <(G_{\mu\nu}^{a})^2>_{0} +\biggl(1-\frac{2\gamma_{m}(\alpha_s)}{\beta_{0}}\biggr) \sum_{f}<m_{f}\overline{q}_{f}q_{f}>_{0} \label{eq:energymomentum} \end{equation} is renormalization group invariant. On the other hand, in the MS type schemes the quark condensate $<m_{f}\overline{q}_{f}q_{f}>_{0}$ is renormalization group invariant to all orders of perturbation theory (see, e.g., Tarrach, 1982). One can introduce the renormalization group invariant quantity \begin{equation} \Omega = -\frac{\beta(\alpha_s)}{\beta_{0}}<(G_{\mu\nu}^{a})^2>_{0} -\frac{4\gamma_{m}(\alpha_s)}{\beta_{0}} \sum_{f}<m_{f}\overline{q}_{f}q_{f}>_{0} \label{eq:omega} \end{equation} so that the new coefficient functions defined from equation \begin{displaymath} \hspace{-45mm} C_{G^2}\biggl(\frac{\mu^2}{Q^2},\alpha_s\biggr)<(G_{\mu\nu}^{a})^2>_{0} +C_{\overline{q}q}^{f}\biggl(\frac{\mu^2}{Q^2},\alpha_s\biggr) \sum_{f}<m_{f}\overline{q}_{f}q_{f}>_{0} \end{displaymath} \begin{equation} \hspace{4cm} =\overline{C}_{G^2}\biggl(\frac{\mu^2}{Q^2},\alpha_s\biggr)\Omega +\overline{C}_{\overline{q}q}^{f}\biggl(\frac{\mu^2}{Q^2},\alpha_s\biggr) \sum_{f}<m_{f}\overline{q}_{f}q_{f}>_{0} \label{eq:CFequation} \end{equation} should be the renormalization group invariants. This is true since the l.h.s of eq.\ (\ref{eq:CFequation}) is directly connected to the observables (Shifman, Vainshtein and Zakharov, 1979) and consequently is invariant. {}From eqs.\ (\ref{eq:omega}) and (\ref{eq:CFequation}) we find the invariant coefficient functions corresponding to the invariant combinations of the gluon and quark condensates \begin{equation} \overline{C}_{G^2}\biggl(\frac{\mu^2}{Q^2},\alpha_s\biggr) =-\frac{\beta_{0}}{\beta(\alpha_s)} C_{G^2}\biggl(\frac{\mu^2}{Q^2},\alpha_s\biggr), \label{eq:CFggnew} \end{equation} \begin{equation} \overline{C}_{\overline{q}q}^{f}\biggl(\frac{\mu^2}{Q^2},\alpha_s\biggr) =C_{\overline{q}q}^{f}\biggl(\frac{\mu^2}{Q^2},\alpha_s\biggr) -\frac{4\gamma_{m}(\alpha_s)}{\beta(\alpha_s)} C_{G^2}\biggl(\frac{\mu^2}{Q^2},\alpha_s\biggr). \label{eq:CFqqnew} \end{equation} Note that, in fact, there are terms of the type $m_{f}^2m_{f'}^2$ or/and $m_{f}^4$ in the r.h.s. of eq.\ (\ref{eq:CFequation}). However, obviously these terms do not affect our equations for invariant coefficient functions. The two-loop coefficient functions for $\sim m^4$ terms have been calculated by Chetyrkin, Gorishny and Spiridonov (1985). The contributions from such terms are negligible for phenomenological applications and will not be discussed here. Now one can use the renormalization group invariance of the coefficient functions and write \begin{equation} \overline{C}_{i}\biggl(\frac{\mu^2}{Q^2},\alpha_s\biggr) =\overline{C}_{i}(1,\alpha_s(Q^2)). \label{eq:CFRGnew} \end{equation} Reevaluating the coefficient functions for the $u,d,s$ light quarks ($N=3$) we obtain the following results in the $\overline{\mbox{MS}}$ scheme. \noindent In the vector channel \begin{equation} \overline{C}_{G^2}(\alpha_s(Q^2)) =\frac{1}{Q^4}\frac{1}{12}\biggl(1-\frac{\alpha_s(Q^2)}{\pi} 0.6111+O(\alpha_{s}^{2})\biggr), \label{eq:CggnewV} \end{equation} \begin{equation} \overline{C}_{\overline{q}q}^{f=f'}(\alpha_s(Q^2)) =\frac{1}{Q^4}2\biggl[1+0.4074\frac{\alpha_s(Q^2)}{\pi} \biggl(1+\frac{\alpha_s(Q^2)}{\pi}14.8180+O(\alpha_{s}^{2})\biggr) \biggr]. \label{eq:CqqnewV} \end{equation} In the scalar channel \begin{equation} \overline{C}_{G^2}(\alpha_s(Q^2)) =\frac{1}{Q^2}\frac{1}{8}\biggl(1+\frac{\alpha_s(Q^2)}{\pi} 3.7222+O(\alpha_{s}^{2})\biggr), \label{eq:CggnewS} \end{equation} \begin{equation} \overline{C}_{\overline{q}q}^{f=f'}(\alpha_s(Q^2)) =\frac{1}{Q^2}3\biggl[1+4.4074\frac{\alpha_s(Q^2)}{\pi} \biggl(1+\frac{\alpha_s(Q^2)}{\pi}7.6879+O(\alpha_{s}^{2})\biggr) \biggr]. \label{eq:CqqnewS} \end{equation} In the pseudoscalar channel \begin{equation} \overline{C}_{\overline{q}q}^{f=f'}(\alpha_s(Q^2)) =-\frac{1}{Q^2}\biggl[1+5.4444\frac{\alpha_s(Q^2)}{\pi} \biggl(1+\frac{\alpha_s(Q^2)}{\pi}9.4559+O(\alpha_{s}^{2})\biggr) \biggr]. \label{eq:CqqnewPS} \end{equation} For all channels \begin{equation} \overline{C}_{\overline{q}q}^{f \neq f'}(\alpha_s(Q^2)) =C_{\overline{q}q}^{f \neq f'}(\alpha_s(Q^2)) +O(\alpha_{s}^3). \label{eq:Cqqffnew} \end{equation} Note again very large $O(\alpha_{s}^2)$ corrections for the coefficient functions of quark condensates in the $\overline{\mbox{MS}}$-scheme. The running coupling is evaluated at the typical hadronic mass scale. Presently the $O(\alpha_{s})$ corrections have also been calculated for the $\mbox{dim}=6$ operators (Lanin, Spiridonov, and Chetyrkin (1986)). We also mention the calculations in the case of heavy quark currents (see, e.g., Broadhurst et al, 1994 and references therein). \vspace{2cm} \renewcommand{\thesection}{\Roman{section}} \section{\bf $R(s)$ in electron-positron annihilation to $O(\alpha^{3}_{s})$} \renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0} In this section we present an outline of the evaluation of the corrections up to $O(\alpha^{3}_{s})$ to the total cross section in the process $e^{+}e^{-}\rightarrow \mbox{hadrons}$ (Fig.\ 5) in the limit of zero light quark masses and infinitely large top mass. We also mention the QCD evaluation of the hadronic decay rates of the Z boson and the relevant quark mass effects. \vspace{6cm} \begin{center} FIG.\ 5. \hspace{2mm} The process $e^{+}e^{-} \rightarrow \mbox{hadrons}$. The shaded bulb includes any interactions of quarks and gluons (or ghosts) allowed in QCD. The dots cover any relevant number of gluon and quark propagators. \end{center} These calculations were first attempted by Gorishny, Kataev and Larin (1988). However, it was shown that those results were incorrect. Indeed, about 4 years ago an independent calculation of the above quantity was completed (Surguladze and Samuel, 1991a,b). The result is much smaller and has the opposite sign compared with the old 1988 result. This finding was confirmed shortly after that by Gorishny, Kataev and Larin (1991). In the process shown in Fig.\ 5 an electron-positron pair annihilates producing either a photon or a $Z$-boson, which further produces quark-antiquark pairs (in QED) plus gluons (if strong interactions are ``switched on''). Finally, quarks through hadronization form hadronic final states with probability equal to one (confinement hypothesis) and the total cross-section is given by \begin{equation} \sigma_{\mbox{\scriptsize{tot}}}(e^{+}e^{-}\rightarrow \mbox{hadrons})=\frac{4\pi\alpha^2}{3s} 3\sum_{f}Q_{f}^2(1+\delta_{\mbox{\tiny QCD}}), \label{eq:sigmatot} \end{equation} where $s$ is the total centre-of-mass energy squared, $Q_f$ is the electric charge of the participating at the given energy quark flavor $f$, factor 3 stands for the number of color degrees of freedom and $\delta_{\mbox{\tiny QCD}}$ stands for the strong interaction contributions. The hadronic production in electron-positron annihilation is usually characterized in terms of the $R$-ratio -the total hadronic cross section normalized by the muon pair production cross section \begin{equation} R(s) = \frac{\sigma_{\mbox{\scriptsize{tot}}}(e^{+}e^{-}\rightarrow \mbox{hadrons})} {\sigma(e^{+}e^{-}\rightarrow \mu^{+}\mu^{-})} = 3\sum_{f}Q_{f}^2(1+\delta_{\mbox{\tiny{QCD}}}). \label{eq:Rratio} \end{equation} The above expressions are relevant at energies much less than the Z mass ($\sqrt{s} \ll M_Z$) corresponding to, for instance, PEP/PETRA energy range. At LEP the effects of the Z boson become important. The corresponding $R$-ratio is defined as a ratio of the hadronic and electronic widths of the Z boson \begin{equation} R_{Z} = \frac{\Gamma(Z \rightarrow \mbox{hadrons})} {\Gamma(Z \rightarrow e^{+}e^{-})}. \label{eq:Rzratio} \end{equation} Note that the total hadronic width of the Z boson in the above equation is the sum of the vector and axial current induced decay rates. Strictly speaking, those rates get different strong interaction contributions. In the present section we calculate the QCD corrections in the vector channel - $\delta_{\mbox{\tiny QCD}}$ in the limit of massless light quarks and the infinitely large top mass. This quantity is, in fact, relevant for the axial part as well. To get the complete axial decay rate, additional contributions are necessary. For details see the original works: Kniehl and K\"{u}hn (1990), Kniehl (1990), Chetyrkin and K\"{u}hn (1990), Chetyrkin, K\"{u}hn and Kwiatkowski (1992), Chetyrkin (1993a), Soper and Surguladze (1994), Surguladze (1994c), the review articles by Kniehl (1994b, 1995b), Soper and Surguladze (1995) and also section 4 of the present paper. \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf $R(s)$ via renormalization constants} The vacuum polarization function $\Pi(Q^2)$ defined in eq.\ (\ref{eq:pifunction}) has a cut along the negative $Q^2$ axis in the massless case. The ratio $R(s)$ can be found taking the imaginary part of $\Pi(s+i0)$, according to eq.\ (\ref{eq:Rsdefin}). Alternatively, $R(s)$ can also be found from eq.\ (\ref{eq:Ddefin}), which in combination with eq.\ (\ref{eq:RGDsolut}) gives to $O(\alpha_s^3)$ \begin{equation} R(s)=R_0+\frac{\alpha_s(s)}{\pi}R_1 +\biggl(\frac{\alpha_s(s)}{\pi}\biggr)^2 R_2 +\biggl(\frac{\alpha_s(s)}{\pi}\biggr)^3 \biggl(R_3-\frac{\pi^2\beta_0^2}{3}R_{1}\biggr). \label{eq:Rexpans} \end{equation} The origin of the large and negative scheme-scale independent term $R_{1}\pi^2\beta_0^2/3$ can be understood if one takes into account the presence of $\sim \log^3\mu^2/s$ terms at $O(\alpha_s^3)$ in the $\Pi$-function and \begin{displaymath} \frac{1}{\pi}\mbox{Im}\log^3(s+i0)=-3\log^2 s+\pi^2. \end{displaymath} The leading QCD term $R_1$ at $O(\alpha_s^3)$ is due to the coupling renormalization. Note, that the $R_i$ in the above equation are the perturbative coefficients of the $D(Q^2)$ function defined in eq.\ (\ref{eq:Ddefin}). For the discussion of the procedure of analytical continuation and the origin of additional $\sim \pi^2$ terms, see also Krasnikov and Pivovarov (1982), Pennington and Ross (1982), Radyushkin (1982), and Pivovarov (1992a). Substituting eq.\ (\ref{eq:Piexpans}) with the renormalized strong coupling into eq.\ (\ref{eq:Rsdefin}) and taking into account the relations (\ref{eq:Pirelat0}), (\ref{eq:Pirelat1}), (\ref{eq:Pirelat2}), (\ref{eq:Pirelat3}) we obtain \begin{eqnarray} \lefteqn{\hspace{-9mm}R(s) = -\frac{3}{4}\biggl\{ Z_{1,-1}+\frac{\alpha_s(\mu)}{\pi}(2Z_{2,-1}) +\left(\frac{\alpha_s(\mu)}{\pi}\right)^2 \biggl(3Z_{3,-1}-\beta_0\Pi_{2,0} +2\beta_0Z_{2,-1}\log\frac{\mu^2}{s}\biggr) } \nonumber\\ && \quad \hspace{-7mm} +\left(\frac{\alpha_s(\mu)}{\pi}\right)^3 \biggl[4Z_{4,-1}-2\beta_0\Pi_{3,0}-\beta_1\Pi_{2,0} +2\beta_0^2\Pi_{2,1}-\frac{2\pi^2\beta_0^2}{3}Z_{2,-1} \nonumber\\ && \quad +(6\beta_0Z_{3,-1}+2\beta_1Z_{2,-1}-2\beta_0^2\Pi_{2,0})\log\frac{\mu^2}{s} +2\beta_0^2Z_{2,-1}\log^2\frac{\mu^2}{s}\biggr]+O(\alpha_s^4)\biggr\}. \label{eq:Rmain} \end{eqnarray} Note that the appearance of perturbative coefficients of the renormalization constant in the above equation is totally due to the relations (\ref{eq:Pirelat3}). In fact, $Z_{\Pi}$ has only simple poles and hence has no imaginary part. The latter is the specific feature of the MS prescription. The expression (\ref{eq:Rmain}) exhibits one of the main ideas of this calculation. Namely, in order to calculate the $l$-loop contribution to $R$, it suffices to calculate the $l$-loop counterterm $Z_{\Pi}$ to the bare quantity $\Pi^{\mbox{\scriptsize B}}$, and the $l-1$ -loop approximation to $\Pi^{\mbox{\scriptsize B}}$. In other words, the minimal information necessary to obtain the four-loop $R(s)$ is contained in the divergent part of one-loop diagram, two-loop diagrams calculated up to $\sim \varepsilon$ terms, three-loop diagrams calculated up to the finite parts in the limit $\varepsilon \rightarrow 0$ and only a leading $\sim 1/\varepsilon$ terms in the overall counterterms of the four-loop diagrams. In fact, as we demonstrate in the next subsection, using the infrared rearrangement procedure (Vladimirov, 1980; Chetyrkin and Tkachov, 1982) one can complete the entire calculation dealing effectively only with three-loop diagrams. We mention once again that, through the procedure of infrared rearangement, within the MS prescription, the problem of calculation of the counterterms to arbitrary $l$ -loop diagrams with an arbitrary number of masses and external momenta can be reduced to the calculation of $l-1$ -loop propagator type massless integrals up to finite terms. In our case $l=4$. On the other hand, the recursive type algorithms for multiloop Feynman integrals (Chetyrkin and Tkachov, 1981; Tkachov, 1981, 1983) and their computer implementation (Surguladze and Tkachov, 1989; Surguladze, 1989b,c, 1992; Gorishny, Larin, Surguladze and Tkachov, 1989) allow one to calculate propagator type Feynman diagrams to three-loop level. \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf Full calculational procedure with a typical four-loop diagram} In this subsection we demonstrate the full calculational procedure for a typical four-loop diagram pictured in Fig.\ 6, which contributes to the photon renormalization constant $Z_{\Pi}$ and hence to the $R$-ratio. To simplify the description, in some cases we will avoid complicated equations, substituting for them their graphical representation. \newpage \mbox{} \vspace{33mm} \begin{center} FIG.\ 6. \hspace{2mm} A typical four-loop nonplanar type diagram contributing to $R(s)$ \end{center} We need to evaluate the counterterm to the diagram pictured in Fig.\ 6. In other words, we should evaluate $-{\cal K}R'$ for this diagram. A simple power counting shows that the given diagram diverges as \begin{displaymath} G\sim\lim_{Q \rightarrow \infty}Q^{4D-14} \end{displaymath} and the superficial degree of divergence is 2. Using the fact that the counterterm has only a polynomial dependence on the external momenta $Q$ within the MS prescription, one can remove such a dependence by differentiating the diagram twice with respect to $Q$ and then set the external momentum to zero. At the next step, since there is no dependence on the external momentum, one can introduce a new fictitious external momentum flowing through one of the diagram lines. This line should be chosen in a way that simplifies the topology of the diagram and avoids infrared divergences. The above procedure for the diagram in Fig.\ 6 is displayed in the following graphical equation \vspace{1cm} \begin{displaymath} Z\supset{\cal K}R'\biggl\{ \biggl(\frac{\partial}{\partial Q_{\mu}}\biggr)^2 \hspace{43mm} \biggr\}_{Q=0} \end{displaymath} \vspace{2cm} \begin{equation} \sim{\cal K}R'\biggl\{4(2-D)\hspace{37mm}+2\hspace{37mm}\biggr\} \label{eq:IR} \end{equation} \vspace{2cm} \begin{displaymath} =\hspace{5mm}{\cal K}\biggl\{4(2-D)\hspace{36mm}+2\hspace{36mm}\biggr\}, \end{displaymath} \vspace{2cm} \noindent where the dot and dashes on the lines result from differentiating the corresponding fermion propagators \begin{displaymath} \biggl(\frac{\partial}{\partial Q_{\mu}}\biggr)^2 \biggl[-\hspace{-2mm}-\hspace{-2mm} \stackrel{P+Q}{\longleftarrow}\hspace{-2mm}-\hspace{-2mm}- \biggr]_{Q=0} \equiv 2(2-D) \biggl[\longleftarrow \hspace{-2mm} \stackrel{P}{\bullet}\hspace{-2mm}\longleftarrow \biggr]=2(2-D)\frac{\hat{P}}{P^4}, \end{displaymath} \begin{displaymath} \frac{\partial}{\partial Q_{\mu}} \biggl[-\hspace{-2mm}-\hspace{-2mm} \stackrel{P+Q}{\longleftarrow}\hspace{-2mm}-\hspace{-2mm}- \biggr]_{Q=0}\equiv \biggl[\longleftarrow \hspace{-2mm} \hspace{-2mm}\stackrel{P}{\setminus}\longleftarrow\biggr] =-\frac{\hat{P}}{P^2}\gamma^{\mu}\frac{\hat{P}}{P^2} \end{displaymath} Boxes contain the corresponding three-loop propagator type subgraphs with subtracted divergences - complete $R$-operation (Fig.\ 7). The dotted lines mean that this line is temporarily ``torn''. After the evaluation of boxes, the parts of the torn line should be pasted and a trivial fourth loop integration should be done, taking into account the corresponding exponents of the propagators due to the three-, two- and one-loop ``box'' insertions. The above procedure gives a great simplification of the problem. Indeed, the evaluation of the four-loop counterterm is reduced to the evaluation of three-, two- and one-loop graphs. \vspace{15mm} \begin{displaymath} -\fbox{{\bf A}}-\equiv R \biggl\{ \hspace{19mm} \biggr\} = \hspace{19mm}-\biggl(\hspace{17mm}\biggr)\hspace{19mm} -\biggl(\hspace{17mm}\biggr) \hspace{17mm} \end{displaymath} \vspace{2cm} \begin{displaymath} \hspace{-43mm}-\fbox{{\bf B}}-\equiv R \biggl\{ \hspace{19mm} \biggr\} = \hspace{19mm}-\biggl(\hspace{17mm}\biggr)\hspace{19mm} \end{displaymath} \vspace{1cm} \begin{center} FIG.\ 7. \hspace{2mm} Complete $R$-operation for the three-loop subgraphs \end{center} The complete $R$-operation of the three-loop diagram insertions corresponding to the ones at the r.h.s. of eq.\ (\ref{eq:IR}) is given in Fig.\ 7. Graphs in the brackets correspond to two- and three-loop counterterms. There is no one-loop divergent subgraph in this particular diagram. Thus, \begin{displaymath} \biggl( G_{i} \biggr) \equiv {\cal K}R' \biggl\{ G_{i} \biggr\}, \end{displaymath} where $G_{i}$ is any divergent subgraph of the given diagram. It is easy to recognize that the two-loop subgraph in Fig.\ 7 does not have subdivergences (only an overall one) and the corresponding counterterm is simply the pole part of this subgraph \vspace{7mm} \begin{displaymath} {\cal K}R' \biggl\{ \hspace{3cm} \biggr\} = {\cal K}\biggl\{ \hspace{3cm} \biggr\}. \end{displaymath} \vspace{7mm} Analogously, because of the topology, the three-loop counterterm does not have a subdivergence and the corresponding counterterm is the pole part of this diagram \vspace{7mm} \begin{displaymath} {\cal K}R' \biggl\{ \hspace{3cm} \biggr\} = {\cal K}\biggl\{ \hspace{3cm} \biggr\}. \end{displaymath} \vspace{7mm} If, in general, a diagram contains divergent subgraphs, then the recursive formula (\ref{eq:KR'}) should be used. As a result of the above manipulations, we managed to reduce the problem of calculation of the counterterm to the four-loop diagram pictured in Fig.\ 6 to the calculation of several three-, two- and one-loop diagrams shown in Fig.\ 7. Note, however that the ``dots'' and ``dashes'' on the diagram lines make their evaluation significantly more difficult. The computer programs for analytical programming systems capable of handling such calculations are the {\small SCHOONSCHIP} program {\small MINCER} (Gorishny, Larin, Surguladze and Tkachov, 1989; Surguladze, 1989b,c) and the {\small FORM} program {\small HEPL}oops (Surguladze, 1992). The latter is especially well-suited for large scale calculations and is much more efficient than the {\small MINCER} program. It is important to stress that, in fact, it is sufficient to evaluate only the ${\cal K}R'$ for the relevant three-loop subgraphs. In other words, it is not necessary to calculate separately three-loop counterterms similar to the graph in the last brackets for the box A in Fig.\ 7. Indeed, a more detailed analysis gives \begin{equation} R[G]=R'[G]-(1-D/2){\cal K}\left(\frac{1}{1-D/2}R'[G]\right), \label{eq:CTrelation} \end{equation} where G is the corresponding three-loop subgraph. The above relation allows simple computer implementation and facilitates calculations considerably. The complete $R$-operation for each four-loop diagram generally has the form \begin{displaymath} \left(\frac{\mu^2}{Q^2}\right)^{4\varepsilon}f_4(\varepsilon) -\sum_{l=1}^{3}\left(\frac{\mu^2}{Q^2}\right)^{(4-l)\varepsilon} c_l(1/\varepsilon)f_{4-l}(\varepsilon), \end{displaymath} where $f_i(\varepsilon)$ is the result of the calculation of the corresponding Feynman graphs including the last trivial loop integration and $c_l$ are the $l$-loop counterterms, polynomials in $1/\varepsilon$. As we already mentioned, in the MS type renormalization scheme, the counterterm for a particular diagram is a polynomial in dimensional parameters (see, e.g., the textbook by Collins, 1984 and references therein). Thus, the terms of the type $(1/\varepsilon)^n\ln^m(\mu^2/Q^2)$, which appear due to the expansion of the factors $(\mu^2/Q^2)^{l\varepsilon}$ into the Laurent series in $\varepsilon$, must be canceled in the final answer for the particular diagram. This can be used to test the calculations at the graph-by-graph level. Recall, that we calculate the counterterm $Z_{4,i}$ to the four-loop diagram. Finally, for the contribution to the $Z_{\Pi}$ of the diagram pictured in Fig.\ 6 we obtain the following result \begin{displaymath} \left(\frac{\alpha_s}{4\pi}\right)^3\hspace{2mm}N_FC_F(C_F-C_A)(C_F-C_A/2) \hspace{2mm} \frac{1}{3-2\varepsilon} \hspace{2mm} \biggl[4\frac{1}{\varepsilon^3}-26\frac{1}{\varepsilon^2} +\frac{65}{4}\frac{1}{\varepsilon}-40\zeta(3)\frac{1}{\varepsilon}\biggr]. \end{displaymath} The {\small CPU} time for the above diagram on a 0.8 MFlop {\small IBM} compatible mainframe was over 6 hours. The extended version of the program {\small MINCER} for the system {\small SCHOONSCHIP} was used. Note that the above result, as well as the total result for the photon renormalization constant does not depend on any modification of the minimal subtraction prescription. \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf Four-loop results} In this subsection, we present results and some of the details of the $O(\alpha_s^3)$ QCD evaluation of the ratio $R(s)$ in electron-positron annihilation (Surguladze and Samuel, 1991a,b). The total number of topologically distinct Feynman diagrams contributing to $Z_{1,i}$ is 1, to $Z_{2,i}$ is 2, to $Z_{3,i}$ is 17 and to $Z_{4,i}$ is 98. However, after application of the infrared rearrangement procedure which, as discussed above, involves differentiation twice with respect to the external momentum of the diagram, the number of four-loop graphs which need to be calculated increases to approximately 250. Furthermore, there are one-, two- and three-loop diagrams, approximately 600, which need to be calculated to subtract subdivergences (evaluate $R'$) for all four-loop diagrams. All analytical calculations of the four-loop diagrams have been done by using the program, which is an extended version (Surguladze, 1989c) of the program {\small MINCER} (Gorishny, Larin, Surguladze and Tkachov, 1989; Surguladze, 1989b). This version includes new subprograms for 4th loop integration and for ultraviolet renormalization. Evaluation of one- and two-loop counterterms has been done by using the program LOOPS (Surguladze and Tkachov, 1989a). The above programs are written on the algebraic programming systems {\small SCHOONSCHIP} (Veltman, 1967; Strubbe, 1974) and {\small REDUCE} (Hearn, 1973) respectively. The full calculation took over 700 hours of {\small CPU} time on three {\small IBM} compatible 0.8 MFlop EC-1037 mainframes with the {\small SCHOONSCHIP} system. We have also recalculated some of the difficult four-loop diagrams with {\small HEPL}oops - a new program for analytical multiloop calculations (Surguladze, 1992). The status of these and some other programs has been reviewed recently in Surguladze (1994d). In the diagram calculations the Feynman gauge is used. The momentum integrations are performed within the $\overline{\mbox{MS}}$ modification (Bardeen, Buras, Duke and Muta, 1978) of the minimal subtraction prescription ('t Hooft, 1973), which amounts to formally setting $\gamma=\zeta(2)=\log4\pi=0$. A discussion of the scheme dependence of the results is given at the end of this section and in section 9. The full graph-by-graph results will be published elsewhere. The analytical result for the four-loop photon renormalization constant reads \begin{eqnarray} \lefteqn{Z_{\mbox{\scriptsize ph}} \equiv 1+\frac{\alpha}{4\pi}Z_{\Pi}=} \nonumber\\ && 1 +N_F\frac{\alpha}{4\pi} \sum_{f}Q_{f}^2\biggl\{ -\frac{4}{3}\frac{1}{\varepsilon} +\frac{\alpha_s}{4\pi}\biggl[\frac{1}{\varepsilon}(-2C_F)\biggr] \nonumber\\ && \quad +\biggl(\frac{\alpha_s}{4\pi}\biggr)^2\biggl[\frac{1}{\varepsilon^2} \biggl(\frac{22}{9}C_FC_A-\frac{8}{9}NTC_F\biggr) +\frac{1}{\varepsilon}\biggl(\frac{2}{3}C_F^2-\frac{133}{27}C_FC_A +\frac{44}{27}NTC_F\biggr)\biggr] \nonumber\\ && \quad +\biggl(\frac{\alpha_s}{4\pi}\biggr)^3 \biggl[\frac{1}{\varepsilon^3} \biggl(-\frac{121}{27}C_FC_A^2+\frac{88}{27}NTC_FC_A -\frac{16}{27}N^2T^2C_F\biggr) \nonumber\\ && \quad +\frac{1}{\varepsilon^2}\biggl(-\frac{11}{9}C_F^2C_A+\frac{2381}{162} C_FC_A^2-\frac{14}{9}NTC_F^2-\frac{778}{81}NTC_FC_A +\frac{88}{81}N^2T^2C_F\biggr) \nonumber\\ && \quad +\frac{1}{\varepsilon}\biggl(\frac{23}{2}C_F^3 +\biggl(-\frac{430}{27}+\frac{88}{9}\zeta(3)\biggr)C_F^2C_A +\biggl(-\frac{5815}{972} -\frac{88}{9}\zeta(3)\biggr)C_FC_A^2 \nonumber\\ && \quad +\biggl(\frac{338}{27}-\frac{176}{9}\zeta(3)\biggr)NTC_F^2 +\biggl(\frac{769}{243} +\frac{176}{9}\zeta(3)\biggr)NTC_FC_A +\frac{308}{243}N^2T^2C_F\biggr)\biggr] \nonumber\\ && \quad + O(\alpha_s^4) \biggr\} +\frac{\alpha}{4\pi} \biggl(\frac{\alpha_s}{4\pi}\biggr)^3\biggl(\sum_{f}Q_f\biggr)^2 \biggl(\frac{d^{abc}}{4}\biggr)^2\biggl(-\frac{176}{9} +\frac{128}{3}\zeta(3)\biggr)\frac{1}{\varepsilon}. \label{eq:Zanalres} \end{eqnarray} It should be stressed that the Riemann $\zeta$-functions $\zeta(4)$ and $\zeta(5)$, which appear at the individual graph level cancel in the above expression. Moreover, as we have observed, the $\zeta(4)$ has disappeared within each gauge invariant set of diagrams. Note that $\zeta(3)$ disappears for QED ($C_F=1, C_A=0, T=1$) except the last term, which comes from the ``light-by-light'' type diagrams (Fig.\ 8). The diagrams pictured in Fig.\ 8 are some of the most complicated ones and the computation of each of them requires over 80h of {\small CPU} time. Note, however, that the second and fourth diagrams in Fig.\ 8 differ correspondingly from the first and third ones only by the SU(N) group weights. So, in fact, only two of them have been calculated. The result (\ref{eq:Zanalres}) does not depend on the particular modification of the minimal subtraction prescription. \vspace{35mm} \begin{center} FIG.\ 8. \hspace{2mm} ``Light-by-light'' type diagrams \end{center} \vspace{5mm} In order to evaluate $R(s)$ to $O(\alpha_s^3)$, besides the four-loop $Z_{\Pi}$ we calculate the unrenormalized hadronic vacuum polarization function $\Pi^{\mbox{\scriptsize B}}(Q^2)$ to the three-loop level. We get the following analytical result in the $\overline{\mbox{MS}}$ scheme. \newpage \begin{eqnarray} \lefteqn{\Pi^{\mbox{\scriptsize B}} \biggl(\frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{Q^2} ,\alpha_{s}^{\mbox{\tiny B}}\biggr)=} \nonumber\\ && N_F\sum_{f}Q_{f}^2\biggl\{ \biggl(\frac{\mu_{\overline{\mbox{\tiny MS}}}^2} {Q^2}\biggr)^\varepsilon \biggl[ \frac{4}{3}\frac{1}{\varepsilon} +\frac{20}{9}+\frac{112}{27}\varepsilon +\frac{656}{81}\varepsilon^2 -\frac{28}{9}\zeta(3)\varepsilon^2 \biggr] \nonumber\\ && \quad +\biggl(\frac{\alpha_{s}^{\mbox{\tiny B}}}{4\pi}\biggr) \biggl(\frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{Q^2}\biggr)^{2\varepsilon} C_{F}\biggl[ 2\frac{1}{\varepsilon} +\frac{55}{3}-16\zeta(3) +\varepsilon \biggl(\frac{1711}{18}-\frac{152}{3}\zeta(3) -24\zeta(4) \biggr) \biggr] \nonumber\\ && \quad +\biggl(\frac{\alpha_{s}^{\mbox{\tiny B}}} {4\pi}\biggr)^2 \biggl(\frac{\mu_{\overline{\mbox{\tiny MS}}}^2}{Q^2}\biggr)^{3\varepsilon} \biggl[ C_{F}^{2}\biggl( -\frac{2}{3}\frac{1}{\varepsilon}-\frac{286}{9} -\frac{296}{3}\zeta(3)+160\zeta(5) \biggr) \nonumber\\ && \quad +C_{F}C_{A} \biggl(\frac{44}{9}\frac{1}{\varepsilon^2} +\frac{1948}{27}\frac{1}{\varepsilon} -\frac{176}{3}\zeta(3)\frac{1}{\varepsilon}+\frac{50339}{81} -\frac{3488}{9}\zeta(3)-88\zeta(4) -\frac{80}{3}\zeta(5) \biggr) \nonumber\\ && \quad +NTC_F \biggl( -\frac{16}{9}\frac{1}{\varepsilon^2} -\frac{704}{27}\frac{1}{\varepsilon} +\frac{64}{3}\zeta(3)\frac{1}{\varepsilon}-\frac{17668}{81} +\frac{1216}{9}\zeta(3)+32\zeta(4) \biggr) \biggr] \biggr\}. \label{eq:Pi3lres} \end{eqnarray} The above result depends on the particular modifications of the minimal subtraction prescription, unlike the result for the renormalization constant (\ref{eq:Zanalres}). Substituting the expressions for the relevant $Z_{i,j}$ and $\Pi_{i,j}$, extracted by comparing eqs.\ (\ref{eq:Zanalres}) and (\ref{eq:Pi3lres}) to eqs.\ (\ref{eq:Zexpans}) and (\ref{eq:Piexpans}), into eq.\ (\ref{eq:Rmain}) and recalling the values for $\beta_0$ and $\beta_1$ from eq.\ (\ref{eq:beta}) we get the following $\overline{\mbox{MS}}$ analytical result for $R(s)$ at the four-loop level \begin{eqnarray} \lefteqn{R^{\overline{\mbox{\scriptsize MS}}}(s)=} \nonumber\\ && N_F\sum_{f}Q_{f}^2\biggl\{ 1 +\biggl(\frac{\alpha_s(s)}{4\pi}\biggr)(3C_F) \nonumber\\ && \quad +\biggl(\frac{\alpha_s(s)}{4\pi}\biggr)^2 \biggl[C_F^2\biggl(-\frac{3}{2}\biggr) +C_FC_A\biggl(\frac{123}{2}-44\zeta(3)\biggr) +NTC_F(-22+16\zeta(3))\biggr] \nonumber\\ && \quad +\biggl(\frac{\alpha_s(s)}{4\pi}\biggr)^3 \biggl[C_F^3\biggl(-\frac{69}{2}\biggr) +C_F^2C_A(-127-572\zeta(3)+880\zeta(5)) \nonumber\\ && \quad \hspace{34mm} +C_FC_A^2\biggl(\frac{90445}{54}-\frac{10948}{9}\zeta(3) -\frac{440}{3}\zeta(5)\biggr) \nonumber\\ && \quad \hspace{34mm} +NTC_F^2(-29+304\zeta(3)-320\zeta(5)) \nonumber\\ && \quad \hspace{34mm} +NTC_FC_A\biggl(-\frac{31040}{27}+\frac{7168}{9}\zeta(3) +\frac{160}{3}\zeta(5)\biggr) \nonumber\\ && \quad \hspace{15mm} +N^2T^2C_F\biggl(\frac{4832}{27} -\frac{1216}{9}\zeta(3)\biggr) -\pi^2C_F\biggl(\frac{11}{3}C_A -\frac{4}{3}NT\biggr)^2\biggr]+O(\alpha_s^4)\biggr\} \nonumber\\ && \quad +\biggl(\frac{\alpha_s(s)}{4\pi}\biggr)^3 \biggl(\sum_{f}Q_f\biggr)^2\biggl(\frac{d_{abc}}{4}\biggr)^2 \biggl[\frac{176}{3}-128\zeta(3)\biggr] +O(\alpha_s^4). \label{eq:Ranalytic0} \end{eqnarray} The logarithmic contributions are absorbed in the running coupling by taking $\mu^2=s$. Those contributions will be presented explicitly in section 9. Note that $\zeta(5)$ appears in the final result due to the contributions from $\Pi_{3,0}$. The last term $\sim (\sum_{f}Q_f)^2$ comes from the so called ``light-by-light'' type diagrams (Fig.\ 8). For standard QCD with the SU$_{c}$(3) gauge group we obtain \begin{eqnarray} \lefteqn{\hspace{-26mm} R^{\overline{\mbox{\scriptsize MS}}}(s)=3\sum_{f}Q_{f}^2\biggl\{ 1 +\frac{\alpha_s(s)}{\pi} +\biggl(\frac{\alpha_s(s)}{\pi}\biggr)^2 \biggl[\frac{365}{24}-11\zeta(3)-N\biggl(\frac{11}{12} -\frac{2}{3}\zeta(3)\biggr)\biggr]} \nonumber\\ && \quad +\biggl(\frac{\alpha_s(s)}{\pi}\biggr)^3 \biggl[\frac{87029}{288}-\frac{121}{8}\zeta(2) -\frac{1103}{4}\zeta(3)+\frac{275}{6}\zeta(5) \nonumber\\ && \quad \hspace{19mm} +N\biggl(-\frac{7847}{216}+\frac{11}{6}\zeta(2) +\frac{262}{9}\zeta(3)-\frac{25}{9}\zeta(5)\biggr) \nonumber\\ && \quad \hspace{19mm} +N^2\biggl(\frac{151}{162}-\frac{1}{18}\zeta(2) -\frac{19}{27}\zeta(3)\biggr)\biggr] \biggr\} \nonumber\\ && \quad \hspace{-18mm} +\biggl(\sum_{f}Q_f\biggr)^2 \biggl(\frac{\alpha_s(s)}{\pi}\biggr)^3 \biggl[\frac{55}{72}-\frac{5}{3}\zeta(3)\biggr] +O(\alpha_s^4). \label{eq:Ranalytic1} \end{eqnarray} Finally, taking into account the values for the relevant Riemann $\zeta$ -functions, $\zeta(2)=\pi^2/6$, $\zeta(3)=1.2020569...$ and $\zeta(5)=1.0369278...$ we obtain the numerical form \begin{eqnarray} \lefteqn{R^{\overline{\mbox{\scriptsize MS}}}(s) =3\sum_{f}Q_{f}^2\biggl[1+\frac{\alpha_s(s)}{\pi} +\biggl(\frac{\alpha_s(s)}{\pi}\biggr)^2(1.9857-0.1153N)} \nonumber\\ && \quad \hspace{27mm} +\biggl(\frac{\alpha_s(s)}{\pi}\biggr)^3 (-6.6368-1.2001N-0.0052N^2)\biggr] \nonumber\\ && \quad \hspace{8mm} -\biggl(\sum_{f}Q_f\biggr)^2 \biggl(\frac{\alpha_s(s)}{\pi}\biggr)^3 1.2395 +O(\alpha_s^4). \label{eq:Rnumerical} \end{eqnarray} Note that only 19 four-loop diagrams contribute to the term $\sim N$ and 2 four-loop diagrams contribute to the term $\sim N^2$. The most complicated diagrams are pictured in Fig.\ 9. The {\small CPU} time for each of them was over 100h and the intermediate expression had as many as $\sim$ $10^5-10^6$ terms. \vspace{4cm} \begin{center} FIG.\ 9. \hspace{2mm} Some of the most complicated diagrams \end{center} \vspace{5mm} It is known, that the perturbative coefficients for $R(s)$ are scheme dependent. The above result was obtained in the modified minimal subtraction, the so-called $\overline{\mbox{MS}}$ scheme introduced by Bardeen, Buras, Duke and Muta (1978). While the scheme-scale dependence problem will be discussed in section 9, here we present the results for a couple of other versions of the minimal subtraction scheme. First, we consider the so called G scheme (Chetyrkin and Tkachov, 1979, 1981; Chetyrkin, Kataev and Tkachov, 1980), which is convenient for practical multiloop calculations. The G scheme is defined in such a way that the trivial one-loop integral in this scheme is \begin{displaymath} \mu^{2\varepsilon}\int\frac{d^{4-2\varepsilon}p}{(2\pi)^{4-2\varepsilon}} \frac{1}{p^2(p-k)^2}=\frac{1}{(4\pi)^2} \hspace{1mm} \biggl(\frac{\mu^2}{k^2}\biggr)^{\varepsilon} \hspace{1mm} \frac{1}{\varepsilon}. \end{displaymath} The result for $R(s)$ in this scheme is \begin{eqnarray} \lefteqn{ R^{\mbox{\scriptsize G}}(s)=3\sum_{f}Q_{f}^2\biggl[1+\frac{\alpha_s(s)}{\pi} +\biggl(\frac{\alpha_s(s)}{\pi}\biggr)^2(-3.514+0.218N)} \nonumber\\ && \quad \hspace{26mm} +\biggl(\frac{\alpha_s(s)}{\pi}\biggr)^3 (-10.980-0.692N+0.029N^2)\biggr] \nonumber\\ && \quad \hspace{7mm} -\biggl(\sum_{f}Q_f\biggr)^2 \biggl(\frac{\alpha_s(s)}{\pi}\biggr)^3 1.240 +O(\alpha_s^4). \label{eq:RGnumerical} \end{eqnarray} The parametrization of the running coupling in the above equation has the same form as in eq.\ (\ref{eq:Asparametr}). However, the parameter $\Lambda$ has to be changed to some other parameter $\Lambda_G$. Finally, in the original MS scheme ('t~Hooft, 1973) we get \begin{eqnarray} \lefteqn{ R^{\mbox{\scriptsize MS}}(s)=3\sum_{f}Q_{f}^2\biggl[1+\frac{\alpha_s(s)}{\pi} +\biggl(\frac{\alpha_s(s)}{\pi}\biggr)^2(7.359-0.441N)} \nonumber\\ && \quad \hspace{26mm} +\biggl(\frac{\alpha_s(s)}{\pi}\biggr)^3 (56.026-8.778N+0.176N^2)\biggr] \nonumber\\ && \quad \hspace{7mm} -\biggl(\sum_{f}Q_f\biggr)^2 \biggl(\frac{\alpha_s(s)}{\pi}\biggr)^3 1.240 +O(\alpha_s^4). \label{eq:RMSnumerical} \end{eqnarray} As one can see, starting from $O(\alpha_s^2)$ the results heavily depend on the choice of the particular modifications of the minimal subtraction scheme. This dependence, called renormalization group ambiguity of perturbative results is an important problem and deserves special consideration. We will return to this issue in section 9. Concluding this section, we mention once again that the results of the above described calculation of the four-loop correction to the $R(s)$ have been published in Surguladze and Samuel (1991a,b) and independently \footnote{See however the discussion in the last three paragraphs of section 3 in the review article by Surguladze (1994d).} in Gorishny, Kataev and Larin (1991) and hence, most likely, the above results are reliable. Interesting relations between the radiative corrections for different observables, found by Brodsky and Lu (1994, 1995) serve, in particular, as another confirmation of our results. \vspace{7mm} \renewcommand{\thesection}{\Roman{section}} \section{\bf \hspace{2mm} $\Gamma(\tau^{-}\rightarrow \nu_{\tau}+\mbox{hadrons})$ to $O(\alpha^{3}_{s})$} \renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0} The other important inclusive process for phenomenology and testing the Standard Model is the hadronic decay of the $\tau$ lepton (Fig.\ 10). For a recent review see, for instance, Pich (1994a). For earlier references see Altarelli (1992), Marciano (1992), and Pich (1991). \vspace{57mm} \begin{center} FIG.\ 10. \hspace{2mm} Hadronic decay of the $\tau$-lepton \end{center} \vspace{2mm} In this section, using our result of four-loop calculation of the $\sigma_{\mbox{\scriptsize{tot}}}(e^{+}e^{-} \rightarrow \mbox{hadrons})$ (Surguladze and Samuel, 1991a,b), we evaluate the hadronic decay rate of the $\tau$ lepton to $O(\alpha_{s}^3)$ in perturbative QCD (Pich, 1990; Gorishny, Kataev and Larin, 1991; Samuel and Surguladze, 1991; see also Braaten, Narison and Pich, 1991; Pich, 1992a,b; Diberder and Pich, 1992a,b; Pivovarov, 1992b). We also comment on the status of the nonperturbative corrections to this quantity. We follow the method first suggested by Tsai (1971), Shankar (1977), and Lam and Yan (1977) for theoretical evaluation of heavy lepton decay rates. This method has been further developed for the $\tau$ lepton including the higher order perturbative corrections and involving the operator product expansion technique (Wilson, 1969) to analyze the nonperturbative contributions (Schilcher and Tran, 1984; Braaten, 1988; Narison and Pich, 1988). As was shown in the above works, combining the operator product expansion technique and analyticity properties of the correlation function of quark currents, the ratio \begin{equation} R_{\tau}=\frac{\Gamma(\tau^{-}\rightarrow\nu_{\tau}+\mbox{hadrons})} {\Gamma(\tau^{-}\rightarrow\nu_{\tau}e^{-}\overline{\nu}_{e})} \label{eq:Rtaudef} \end{equation} is calculable in perturbative QCD. Strictly speaking, besides the QCD perturbative parts the nonperturbative and weak contributions should be included to estimate $R_{\tau}$. There are instanton contributions as well. However, it was shown recently by Nason and Porrati (1993) (see also Kartvelishvili and Margvelashvili, 1995) that these contributions are completely negligible due to the chiral suppression factor $m_{u}m_{d}m_{s}/M_{\tau}^2$. The $R_{\tau}$ can be written as the following sum \begin{equation} R_{\tau}=R_{\tau}^{\mbox{\scriptsize{pert}}} +R_{\tau}^{\mbox{\scriptsize{nonpert}}} +R_{\tau}^{\mbox{\scriptsize{weak}}}. \label{eq:Rtausum} \end{equation} \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf Perturbative QCD contributions} The quantity $R_{\tau}^{\mbox{\scriptsize{pert}}}$ can be expressed as the following integral over the invariant mass of the hadronic decay products of the $\tau$ lepton (Lam and Yan, 1977; Braaten, 1988) \begin{equation} R_{\tau}^{\mbox{\scriptsize{pert}}} =\frac{3}{4\pi}\int_{0}^{M_{\tau}^2}\frac{ds}{M_{\tau}^2} \biggl(1-\frac{s}{M_{\tau}^2}\biggr)^2 \biggl[\biggl(1+2\frac{s}{M_{\tau}^2}\biggr)\mbox{Im}\Pi^{T}(s+i0) +\mbox{Im}\Pi^{L}(s+i0)\biggr], \label{eq:IntM} \end{equation} where $M_{\tau}$ is the mass of the $\tau$ lepton. The functions $\Pi^T$ and $\Pi^L$ are the transverse and longitudinal parts of the correlation function of weak currents of quarks coupled to W boson. In fact, $\Pi^{T,L}$ are the appropriate combinations of vector and axial parts corresponding to the vector and axial currents of u, d, s light quarks (for details see, e.g., Pich, 1994a). The expression for $R_{\tau}^{\mbox{\scriptsize{pert}}}$ in the form of (\ref{eq:IntM}) is not quite useful. The problem is that the correlation functions involved can not be calculated at low energies because of the large nonperturbative effects that invalidate perturbative approach. However, simple analyticity properties of the correlation functions allow us to evaluate the integral in (\ref{eq:IntM}). Indeed, the function $\Pi$ is analytic in the complex $s$ plane everywhere except the positive real axis. According to the Cauchy integral theorem, an integral over $s$ along the closed contour $C_1+C_2$ (Fig.\ 11) of the product of $\Pi(s)$ with any nonsingular function $f(s)$ is zero. \vspace{63mm} \begin{center} FIG.\ 11. \hspace{2mm} Integration contour \end{center} On the other hand, the imaginary part of the correlation function is proportional to its discontinuity across the positive real axis. So, the following relation holds \begin{equation} \int_{0}^{M_\tau^2} \hspace{1mm} ds \hspace{1mm} f(s)\mbox{Im}\Pi(s) = \frac{1}{2i} \int_{C_1} \hspace{1mm} ds \hspace{1mm} f(s)\Pi(s) = -\frac{1}{2i} \int_{C_2} \hspace{1mm} ds \hspace{1mm} f(s)\Pi(s), \label{eq:disc} \end{equation} where the $C_2$ is the circle of radius $\mid s\mid =M_{\tau}^2$ (Fig.\ 11). The benefit of the above relation is that in the r.h.s. one needs to calculate the correlation function for $\mid s\mid$ at $M_{\tau}^2$. Hopefully, $M_{\tau}$ is large enough to use the operator product expansion in powers of $1/M_{\tau}^2$ and the $\alpha_s(M_{\tau})$ is small enough to use perturbative expansion in $\alpha_s$. Then the perturbative method can, in principle, be used to calculate the leading term in the operator product expansion and the higher twist terms can be estimated semi-phenomenologically. Using eq.\ (\ref{eq:disc}), the perturbative part of the ratio $R_{\tau}$ can be expressed by an integral over the invariant mass $s$ of the final state hadrons along the contour $C_2$ in the complex $s$-plane (Fig.\ 11). In the chiral limit, $m_{u}=m_{d}=m_{s}=0$, the currents are conserved and the longitudinal part of the $\Pi(s)$ is absent. In the axial channel $\Pi^L(s)=O(m_f^2/s)$ (see section 4). For the $R_{\tau}^{\mbox{\scriptsize{pert}}}$ we get \begin{equation} R_{\tau}^{\mbox{\scriptsize{pert}}}=\frac{3i}{8\pi}\int_{C_2}\frac{ds} {M_{\tau}^2} \biggl(1-\frac{s}{M_{\tau}^2}\biggr)^2 \biggl[\biggl(1+2\frac{s}{M_{\tau}^2}\biggr)\Pi^{T}(s)\biggr]. \label{eq:Contourint} \end{equation} Note that the factor $(1-s/M_{\tau}^2)^2$ suppresses the contribution from the region near the positive real axis where the $\Pi(s)$ has a branch cut (Braaten, 1988). To simplify the description, we use the chiral limit which is a perfect approximation for $R_{\tau}$. On the other hand, the mass corrections can be included with the calculation very similar to that in section 4. The actual calculations show (Chetyrkin and Kwiatkowski, 1993; see also recent analyses in Pich, 1994a) that the effects of quark mass corrections on $R_{\tau}$ are well below 1\% and can be neglected. Note also that, in the massless quark limit the contributions from vector and axial channels to $\Pi$ coincide at any given order of perturbation theory and evidently the results are flavor independent. So, in this case, for evaluation of $\Pi^T(s)$ in eq.\ (\ref{eq:Contourint}) we use our earlier results for the electromagnetic two-point correlation function that contributes to R(s) in electron-positron annihilation (section 6). The function $\Pi^T(s)$ can be related to the $D(s)$ function defined in section 2 as follows \begin{equation} -\frac{3}{4}s\frac{d}{ds}\Pi^T(s)=\frac{\sum_{f=d,s} \mid V_{uf}\mid^2} {\sum_{f}Q_f^2}D(s), \label{eq:Corrfunct} \end{equation} where $V_{ud}$ and $V_{us}$ are the Kobayashi-Maskawa matrix elements. $\mid V_{ud}\mid^2+\mid V_{us}\mid^2=0.998\pm 0.002$ (see, e.g., Pich, 1994b). The factor in the r.h.s of eq.\ (\ref{eq:Corrfunct}) is due to the replacement of the electromagnetic currents by charged weak currents in the correlation function. Note also that evidently the ``light-by-light'' type graphs (Fig.\ 8) do not contribute to the decay width of the $\tau$ lepton. Thus, the term $\sim (\sum_{f}Q_f)^2$ drops out in the $D$ function. The perturbative coefficients of $D(s)$ have been given in the previous section up to the four-loop level in the vector channel (see eqs.\ (\ref{eq:RGDsolut}) and (\ref{eq:Ranalytic1}) ). Performing the contour integration in eq.\ (\ref{eq:Contourint}) using the relations (\ref{eq:Corrfunct}) and (\ref{eq:RGDsolut}), and replacing $\alpha_s(s)$ by $\alpha_s(M_{\tau})$ using the evolution equation (\ref{eq:Astransform}), we obtain in terms of perturbative coefficients of $R(s)$ \begin{eqnarray} \lefteqn{R_{\tau}^{\mbox{\scriptsize{pert}}}= \frac{\mid V_{ud}\mid^2+\mid V_{us}\mid^2}{\sum_{f}Q_{f}^{2}} \biggl\{R_{0} +\frac{\alpha_s(M_{\tau}^2)}{\pi}R_1 +\biggl(\frac{\alpha_s(M_{\tau}^2)}{\pi}\biggr)^2 \biggl(R_{2}+\frac{19}{12}\beta_0 R_1\biggr)} \nonumber\\ && \quad +\biggl(\frac{\alpha_s(M_{\tau}^2)}{\pi}\biggr)^3 \biggl[R_3+\frac{19}{6}R_2\beta_0+\frac{19}{12}R_1\beta_1 +\biggl(\frac{265}{72}-\frac{\pi^2}{3}\biggr)R_1\beta_0^2\biggr] +O(\alpha_s^4)\biggr\}, \label{eq:Rtaumain} \end{eqnarray} where, as we have already mentioned, the term $\sim (\sum_{f}Q_f)^2$ should be omitted in $R_3$. Substituting the relevant expressions for $R_{i}$ and $\beta_{i}$ from the previous sections, we obtain the $O(\alpha_s^3)$ analytical result in the $\overline{\mbox{MS}}$ scheme \begin{eqnarray} \lefteqn{R_{\tau}^{\mbox{\scriptsize{pert}}}(M_{\tau}^2)= N_F(\mid V_{ud}\mid^2+\mid V_{us}\mid^2)\biggl\{ 1 +\frac{\alpha_s(M_{\tau}^2)}{\pi} \biggl(\frac{3}{4}C_F\biggr)} \nonumber\\ && \quad +\biggl(\frac{\alpha_s(M_{\tau}^2)}{\pi}\biggr)^2 \biggl[C_F^2\biggl(-\frac{3}{32}\biggr) +C_FC_A\biggl(\frac{947}{192}-\frac{11}{4}\zeta(3)\biggr) +NTC_F\biggl(-\frac{85}{48}+\zeta(3)\biggr)\biggr] \nonumber\\ && \quad +\biggl(\frac{\alpha_s(M_{\tau}^2)}{\pi}\biggr)^3 \biggl[C_F^3\biggl(-\frac{69}{128}\biggr) +C_F^2C_A\biggl(-\frac{1733}{768}-\frac{143}{16}\zeta(3) +\frac{55}{4}\zeta(5)\biggr) \nonumber\\ && \quad \hspace{34mm} +C_FC_A^2\biggl(\frac{559715}{13824}-\frac{2591}{96}\zeta(3) -\frac{55}{24}\zeta(5)\biggr) \nonumber\\ && \quad \hspace{34mm} +NTC_F^2\biggl(-\frac{125}{192}+\frac{19}{4}\zeta(3)-5\zeta(5)\biggr) \nonumber\\ && \quad \hspace{34mm} +NTC_FC_A\biggl(-\frac{24359}{864}+\frac{73}{4}\zeta(3) +\frac{5}{6}\zeta(5)\biggr) \nonumber\\ && \quad \hspace{5mm} +N^2T^2C_F\biggl(\frac{3935}{864} -\frac{19}{6}\zeta(3)\biggr) -\frac{\pi^2}{64} C_F\biggl(\frac{11}{3}C_A -\frac{4}{3}NT\biggr)^2\biggr]+O(\alpha_s^4)\biggr\}. \label{eq:Rtauanalytic0} \end{eqnarray} Within the standard QCD with the SU$_{c}$(3) gauge group we obtain \begin{eqnarray} \lefteqn{\hspace{-9mm} R_{\tau}^{\mbox{\scriptsize{pert}}}(M_{\tau}^2)= 3(0.998\pm 0.002)\biggl\{ 1 +\frac{\alpha_s(M_{\tau}^2)}{\pi}} \nonumber\\ && \quad +\biggl(\frac{\alpha_s(M_{\tau}^2)}{\pi}\biggr)^2 \biggl[\frac{313}{16}-11\zeta(3)-N\left(\frac{85}{72} -\frac{2}{3}\zeta(3)\right)\biggr] \nonumber\\ && \quad +\biggl(\frac{\alpha_s(M_{\tau}^2)}{\pi}\biggr)^3 \biggl[\frac{544379}{1152}-\frac{121}{8}\zeta(2)-\frac{8917}{24}\zeta(3) +\frac{275}{6}\zeta(5) \nonumber\\ && \quad \hspace{17mm} +N\biggl(-\frac{8203}{144}+\frac{11}{6}\zeta(2) +\frac{733}{18}\zeta(3)-\frac{25}{9}\zeta(5)\biggr) \nonumber\\ && \quad \hspace{17mm} +N^2\biggl(\frac{3935}{2592}-\frac{1}{18}\zeta(2) -\frac{19}{18}\zeta(3)\biggr) \biggr] +O(\alpha_s^4) \biggr\}_{N=3}, \label{eq:Rtauanalytic1} \end{eqnarray} and a numerical form reads \begin{eqnarray} \lefteqn{\hspace{-9mm}R_{\tau}^{\mbox{\scriptsize{pert}}}(M_{\tau}^2)=} \nonumber\\ && \quad \hspace{-21mm} 3(0.998\pm 0.002)\biggl[ 1 +\frac{\alpha_s(M_{\tau})}{\pi} +5.2023 \left(\frac{\alpha_s(M_{\tau})}{\pi}\right)^2 +26.366 \left(\frac{\alpha_s(M_{\tau})}{\pi}\right)^3 +O(\alpha_s^4) \biggr] \label{eq:Rtaunumer} \end{eqnarray} \vspace{5mm} \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf On the Nonperturbative and Electroweak contributions} \indent The nonperturbative contributions to $R_{\tau}$ can be expressed as a power series of corrections in $1/M_{\tau}^2$ \begin{equation} R_{\tau}^{\mbox{\scriptsize{nonpert}}} \sim \frac{C_2{\cal f}(m_f^2(M_{\tau}) ,\theta_c)}{M_{\tau}^2} +\sum_{i\geq2}\frac{C_{2i}<O_{2i}>_0}{M_{\tau}^{2i}}, \label{eq:Rtaunonp} \end{equation} where the $m_f$ are $u, d, s$ running quark masses, $<O_{2i}>_{0}$ are the so-called vacuum condensates, which can be obtained phenomenologically and the $C_i$ are their coefficient functions describing short distance effects. Note that, in eq.\ (\ref{eq:Rtaunonp}) we formally include part of the pure perturbative corrections (the first term) which is due to the nonvanishing $u, d, s$ quark masses. These corrections for the $u$ and $d$ quarks are completely negligible. The contribution coming from the $s$ quark is suppressed by $\sin^2\theta_{C}$ and is also below 1\% (Pich, 1990). Presently, the only way to estimate the strong interaction effects in the condensate contributions is by perturbation theory. The coefficient functions $C_{2i}$ are asymptotic perturbative series in terms of $\alpha_s$. In order to estimate the nonperturbative contributions, one needs to sum up the power series of the QCD perturbative series. In the previous section we have described the calculation of the high-order perturbative QCD contributions to the coefficient functions of the dimension 4 power corrections (gluon, $<\alpha_sG^2>_0$ and quark, $<m_f\overline{q}_fq_f>_0$ condensates). It was shown (Loladze, Surguladze and Tkachov, 1985; Surguladze and Tkachov, 1989b, 1990) that the high-order perturbative corrections to some of the coefficient functions are too large. For instance, for the coefficient function of the condensate $<m_s\overline{s}s>_0$ in the vector channel (see eq.\ (\ref{eq:CqqnewV})) $\Lambda_{\mbox{\scriptsize eff}} \approx 30\Lambda_{\overline{\mbox{\scriptsize MS}}}$. This indicates that the renormalization group invariant criteria to the perturbative calculability of the QCD contributions to the coefficient function is not fulfilled. The coefficient functions of the dimension 6 condensates are calculated up to $O(\alpha_s)$ (Lanin, Spiridonov and Chetyrkin, 1986) and to analyze the corresponding series one needs at least the next to leading correction. The above uncertainty in coefficient functions $C_{2i}$ allows one to estimate the condensate contributions probably not better than their order of magnitude. There is another source of theoretical uncertainties in the evaluation of condensate contributions of dimension 6 and higher, where the operator basis of expansion includes a large number of operators. Presently, there are no precise methods to estimate their matrix elements. For the matrix elements of four quark operators (dimension 6) the vacuum saturation approximation (Shifman, Vainshtein and Zakharov, 1979) is used to express them as the square of the two-quark matrix elements. However, the vacuum saturation approximation is not expected to be precise enough in order to use it in the analyses of the tiny nonperturbative contributions (see, e.g., analysis by Altarelli, 1992; see also a brief discussion in Surguladze and Samuel, 1992b ). Indeed, as it was found by Braaten (1988) and Pich (1990, 1992a,b, 1994a), the nonperturbative corrections are below the $1\%$ level with large theoretical error. The contributions of dim=4 condensates start at $O(\alpha_s^2)$ and thus are suppressed by two powers of $\alpha_s$. The dim=6 and dim=8 corrections are suppressed by the inverse powers of $M_{\tau}$ ($M_{\tau}^6$ and $M_{\tau}^8$ respectively) and are small. On the other hand, the corrections in vector and axial channels have opposite signs and they largely cancel each other, so the total relative error is even larger. In the works by Pumplin (1989, 1990) it was shown that the uncertainty due to threshold effects makes a significant contribution in the theoretical error for $R_{\tau}$. In the works by Altarelli (1992) and Altarelli, Nason and Ridolfi (1994) an ambiguity $\sim \Lambda^2/M_{\tau}^2$ is discussed. Earlier, Zakharov (1992) has argued that such dim=2 terms in eq.\ (\ref{eq:Rtaunonp}) can be generated by ultraviolet renormalons. For an alternative point of view on the effects of possible dim=2 terms, see Narison (1994). However, this issue is still a subject of intensive discussions and likely is far from being settled. Summarizing, we note that the above mentioned major sources of theoretical uncertainties in the evaluation of small power corrections makes certain restriction on the precision theoretical prediction of $R_{\tau}$ and consequently on $\alpha_{s}(M_{\tau})$. Fortunately, the nonperturbative corrections are suppressed and the hadronic decay of the $\tau$ still remains as a good source to extract the low energy $\alpha_s$. Finally, we note that the electroweak contributions $R_{\tau}^{\mbox{\scriptsize{weak}}}$ were calculated by Marciano and Sirlin (1988), and Braaten and Li (1990). Those corrections contain logarithms of $M_{\tau}/M_{Z}$ and are not negligible. The leading order electroweak corrections give roughly $+2\%$ contributions to $R_{\tau}$ (see, e.g., Pich, 1994a). \vspace{2cm} \renewcommand{\thesection}{\Roman{section}} \section{\bf \ \ \ Four-loop QED Renormalization Group Functions} \renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0} In this section we outline the calculation of the standard QED renormalization group functions at the four-loop level in the minimal and momentum subtraction schemes. These quantities can be obtained as an intermediate result of the calculations of $R(s)$, described in the previous sections, by replacing the SU$_{\mbox{\scriptsize c}}$(3) gauge group invariants for the corresponding diagrams in a proper way. The results of two independent calculations of the four-loop QED $\beta$-function by Gorishny, Kataev and Larin (1990), and by Surguladze (1990) have been reported in the joint publications by Gorishny, Kataev, Larin and Surguladze (1991a,c). \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf General formulae} \indent The Lagrangian density of standard QED is \begin{eqnarray} \lefteqn{L_{QED}=} \nonumber\\ && -\frac{1}{4}F_{\mu\nu}F^{\mu\nu} +i\sum_{j}\overline{\psi}_j\gamma^{\mu}D_{\mu}\psi_{j} -\sum_{j}m_{j}\overline{\psi}_j\psi_j -\frac{1}{2\alpha_{\mbox{\tiny{G}}}} \partial_{\mu}A^{\mu}\partial_{\mu}A^{\mu}, \label{eq:QEDlagr} \end{eqnarray} where $F_{\mu\nu}=\partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu}$ and $D_{\mu}=\partial_{\mu}-ieA_{\mu}$. $\alpha_{\mbox{\tiny{G}}}$ is the gauge parameter, $m_j$ are the fermion masses, $\psi$ and $A_{\mu}$ are the fermion and photon fields and $e$ is the electric charge. Renormalization constants are defined by the relations \begin{displaymath} \psi_{\mbox{\tiny{B}}} = \mu^{-\varepsilon}\sqrt{Z_{\mbox{\tiny{F}}}}\psi, \end{displaymath} \begin{equation} A^{\mu}_{\mbox{\tiny{B}}} = \mu^{-\varepsilon}\sqrt{Z_{\mbox{\scriptsize{ph}}}}A^{\mu}, \label{eq:QEDrenormal} \end{equation} \begin{displaymath} \alpha_{\mbox{\tiny{B}}} =\mu^{2\varepsilon} Z_{\alpha} \alpha \hspace{1cm} (\alpha=e^2/4\pi), \end{displaymath} \begin{displaymath} \alpha_{\mbox{\tiny{G}}}^{\mbox{\tiny{B}}} = Z_{\mbox{\tiny{G}}} \alpha_{\mbox{\tiny{G}}}. \end{displaymath} For the fermion-fermion-photon vertex renormalization one has \begin{equation} \mu^{-2\varepsilon}Z_{\mbox{\scriptsize{vert}}} e\overline{\psi}\gamma_{\mu}A^{\mu}\psi =\mu^{-2\varepsilon}\sqrt{Z_{\alpha}Z_{\mbox{\scriptsize{ph}}}} Z_{\mbox{\tiny{F}}}e\overline{\psi}\gamma_{\mu}A^{\mu}\psi. \label{eq:FFPH} \end{equation} According to Ward identity in QED (Ward, 1950) $Z_{\mbox{\scriptsize{vert}}}=Z_{\mbox{\tiny{F}}}$, which implies from eq.\ (\ref{eq:FFPH}) the identity \begin{equation} Z_{\alpha}Z_{\mbox{\scriptsize{ph}}}=1. \label{eq:QEDWI} \end{equation} {}From eqs.\ (\ref{eq:QEDrenormal}) and (\ref{eq:QEDWI}) we get \begin{equation} \alpha_{\mbox{\tiny{B}}} =\mu^{2\varepsilon}Z_{\mbox{\scriptsize{ph}}}^{-1}\alpha. \label{eq:QEDcouplren} \end{equation} The gauge invariance of the QED lagrangian implies the absence of the conterterm for the gauge fixing term in (\ref{eq:QEDlagr}) and, thus, $Z_{\mbox{\tiny{G}}}=Z_{\mbox{\scriptsize{ph}}}$. Using the relation (\ref{eq:QEDcouplren}) and the renormalization group invariance of ``bare'' coupling $\mu^2 d\alpha_{\mbox{\tiny{B}}}/d\mu^2=0$, taking into account that $Z_{\mbox{\scriptsize{ph}}}$ depends on $\mu$ only via $\alpha$ and also the standard definition of the QED MS $\beta$-function \begin{equation} \beta_{\mbox{\tiny{QED}}}^{\mbox{\tiny{MS}}}(\alpha) =\frac{1}{4\pi}\mu^2\frac{d\alpha}{d\mu^2} \biggr|_{\alpha_{\mbox{\tiny{B}}}\mbox{\scriptsize{ fixed}}}, \label{eq:QEDbeta} \end{equation} we obtain a convenient expression for the further evaluation of the $\beta$ function \begin{equation} \beta_{\mbox{\tiny{QED}}}^{\mbox{\tiny{MS}}}(\alpha)= -\frac{1}{4\pi}\lim_{\varepsilon \rightarrow 0} \frac{\varepsilon \alpha}{1-\alpha\frac{\partial}{\partial\alpha} \log Z_{\mbox{\scriptsize{ph}}}}. \label{eq:QEDmain} \end{equation} \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf Four-loop results} \noindent The photon field renormalization constant $Z_{\mbox{\scriptsize{ph}}}$ can be found from the QED relation, analogous to eq.\ (\ref{eq:Zanalres}), where only 58 QED four-loop diagrams contribute to $\Pi(\mu^2/Q^2,\alpha)$. The prescription for the evaluation of the diagram contributions to the $\Pi_{\mbox{\tiny{B}}}$ is analogous to the one described in section 2. The total {\small CPU} time on the three {\small IBM} compatible mainframes was approximately 400 hours. Setting $C_F=1$, $C_A=0$, $T=1$ and $\alpha_s=\alpha$ in eq.\ (\ref{eq:Zanalres}), we obtain the four-loop photon renormalization constant in QED, corresponding to the {\em minimal subtraction} prescription \begin{eqnarray} \lefteqn{\hspace{-1mm}Z_{\mbox{\scriptsize{ph}}}= N-\frac{\alpha}{4\pi}\frac{4}{3\varepsilon}N -\biggl(\frac{\alpha}{4\pi}\biggr)^2 \frac{2}{\varepsilon}N -\biggl(\frac{\alpha}{4\pi}\biggr)^3 \biggl[\frac{8}{9\varepsilon^2}N -\frac{1}{\varepsilon}\biggl(\frac{2}{3}+\frac{44}{27}N\biggr) \biggr]N} \nonumber\\ && \hspace{-4mm} -\biggl(\frac{\alpha}{4\pi}\biggr)^4 \biggl\{\frac{16}{27\varepsilon^3}N^2+\frac{1}{\varepsilon^2} \biggl(\frac{14}{9}N-\frac{88}{81}N^2\biggr) -\frac{1}{\varepsilon} \biggl[\frac{23}{2}-\biggl(\frac{190}{9}-\frac{208}{9}\zeta(3)\biggr)N +\frac{308}{243}N^2 \biggr]\biggr\}N \nonumber\\ \label{eq:ZQED} \end{eqnarray} Substituting the expression for $Z_{\mbox{\scriptsize{ph}}}$ into eq.\ (\ref{eq:QEDmain}), we obtain the following result for the four-loop QED $\beta$-function in the MS type schemes \begin{eqnarray} \lefteqn{\hspace{-23mm}\beta_{\mbox{\tiny{QED}}}^{\mbox{\tiny{MS}}}(\alpha)= \frac{4}{3}N\left(\frac{\alpha}{4\pi}\right)^2 +4N\left(\frac{\alpha}{4\pi}\right)^3 -N\left(2+\frac{44}{9}N\right)\left(\frac{\alpha}{4\pi}\right)^4} \nonumber\\ && \hspace{17mm} -N\biggl[46-\biggl(\frac{760}{27}-\frac{832}{9}\zeta(3)\biggr)N +\frac{1232}{243}N^2\biggr]\biggl(\frac{\alpha}{4\pi}\biggr)^5. \label{eq:betaQED} \end{eqnarray} It is useful for further applications to present the result for the Johnson-Willey-Baker $F_1$ function (Johnson, Willey and Baker, 1967; Baker and Johnson, 1971; Johnson and Baker, 1973). This function can be obtained from the result for $\beta_{\mbox{\tiny{QED}}}^{\mbox{\tiny{MS}}}$ by subtracting the contributions of the diagrams with fermion loop insertions into the photon lines and reducing the power in $\alpha/4\pi$ by one. We obtain \begin{equation} F_1(\alpha)= \frac{4}{3}\left(\frac{\alpha}{4\pi}\right) +4\left(\frac{\alpha}{4\pi}\right)^2 -2\left(\frac{\alpha}{4\pi}\right)^3 -46\left(\frac{\alpha}{4\pi}\right)^4. \label{eq:QEDF1} \end{equation} Note that all coefficients up to four-loop level are rational numbers. The results for most of the individual graphs do contain transcendental $\zeta(3)$, $\zeta(4)$ and $\zeta(5)$. The $\zeta(4)$ and $\zeta(5)$ cancel within each gauge-invariant set of diagrams. The three-loop results agree with the ones obtained by de~Rafael and Rosner (1974). It is possible to recalculate the MS QED $\beta$ function in the form of the Gell-Man-Low $\Psi(\alpha)$ function - the QED $\beta$ function in the MOM scheme. See details in Gorishny, Kataev, Larin and Surguladze (1991a) (see also Adler, 1972; de~Rafael and Rosner, 1974). We obtain the Gell-Mann-Low $\Psi$ function at the four-loop level \begin{eqnarray} \lefteqn{\hspace{-9mm}\Psi(\alpha)= \frac{4}{3}N\biggl(\frac{\alpha}{4\pi}\biggr)^2 +4N\biggl(\frac{\alpha}{4\pi}\biggr)^3 -N\biggl[2 +\biggl(\frac{184}{9}-\frac{64}{3}\zeta(3)\biggr)N\biggr] \biggl(\frac{\alpha}{4\pi}\biggr)^4} \nonumber\\ && -N\biggl[46-\biggl(104+\frac{512}{3}\zeta(3) -\frac{1280}{3}\zeta(5)\biggr)N -\biggl(128-\frac{256}{3}\zeta(3)\biggr)N^2\biggr] \biggl(\frac{\alpha}{4\pi}\biggr)^5. \nonumber\\ \label{eq:QEDpsi} \end{eqnarray} The $O(\alpha^4)$ result agrees with the one obtained by Baker and Johnson (1969) and Acharya and Nigam (1978, 1985). Recently, Broadhurst, Kataev and Tarasov (1993) have carried out an additional calculation necessary to convert the four-loop MS QED $\beta$ function to the four-loop QED on-shell $\beta$ function, usually called the Callan-Symanzik function $\beta_{\mbox{\tiny{QED}}}^{\mbox{\tiny{CS}}}$ (Callan, 1970; Symanzik, 1970, 1971). This function is defined as follows \begin{equation} \beta_{\mbox{\tiny{QED}}}^{\mbox{\tiny{CS}}}(\alpha) =\frac{m_{e}}{\alpha}\frac{d\alpha}{d m_{e}} \biggr|_{\alpha_{\mbox{\tiny{B}}}\mbox{\scriptsize{ fixed}}}, \label{eq:QEDbetaCS} \end{equation} where $m_{e}$ is the electron pole mass. The subtraction prescription in this case requires all subtractions to be on-shell. The three-loop $\beta_{\mbox{\tiny{QED}}}^{\mbox{\tiny{CS}}}$ was calculated long ago by de Rafael and Rosner (1974). The four-loop result has the following form (Broadhurst, Kataev and Tarasov, 1993) \begin{eqnarray} \lefteqn{\beta_{\mbox{\tiny{QED}}}^{\mbox{\tiny{CS}}}(\alpha)= \frac{2}{3}N\left(\frac{\alpha}{\pi}\right) +\frac{1}{2}N\left(\frac{\alpha}{\pi}\right)^2 -N\left(\frac{1}{16}+\frac{7}{9}N\right) \left(\frac{\alpha}{\pi}\right)^3} \nonumber\\ && \hspace{-7mm} -N\biggl[\frac{23}{64}-\biggl(\frac{1}{24}-\frac{5}{3}\zeta(2) +\frac{8}{3}\zeta(2)\ln 2-\frac{35}{48}\zeta(3)\biggr)N -\biggl(\frac{901}{648}-\frac{8}{9}\zeta(2)-\frac{7}{48}\zeta(3)\biggr)N^2 \biggr] \biggl(\frac{\alpha}{\pi}\biggr)^4 \nonumber\\ \label{eq:betaQEDCS} \end{eqnarray} \vspace{2cm} \renewcommand{\thesection}{\Roman{section}} \section{\bf Renormalization Group Ambiguity of Perturbative QCD Predictions} \renewcommand{\thesection}{\arabic{section}} \setcounter{equation}{0} In the previous sections we have demonstrated the calculation of some of the important observables within the framework of perturbative QCD. This involves calculation of a large number of Feynman diagrams and requires a a very large amount of computer and human resources. For example, to $O(\alpha_{s}^{3})$ we have calculated 98 (effectively 250) four-loop Feynman diagrams. The next order requires calculation of approximately 600-700 five-loop diagrams. Calculations of such a scale are extremely difficult. On the other hand, perturbative QCD series are asymptotic ones and the question of how many orders need to be calculated, can be answered only from estimates of remainders (see, e.g., the textbook by Collins, 1984). Moreover, perturbative coefficients beyond the two-loop level, as well as the expansion parameter, are scheme-scale dependent. The scheme-scale ambiguity - a fundamental property of the renormalization group calculations in QCD, does not allow one to obtain reliable estimates from the first few calculated terms without involving additional criteria. In this section we discuss the extraction of reliable estimates for observable quantities within perturbation theory. The problem of scheme-scale dependence of perturbative QCD predictions will be considered first within the MS prescription and then we outline a scheme invariant approach along the lines of Stevenson (1981a,b). We apply the three known approaches for resolving the scheme-scale ambiguity. As a result, we fix the scheme-scale parameter, within the framework of MS prescription, for which all of the criteria tested are satisfied for the quantity $R(s)$ at the four-loop level (Surguladze and Samuel, 1993). On the other hand, we estimate the theoretical error by using the scheme-scale dependence as a measure of the theoretical uncertainty (Surguladze and Samuel, 1993; Surguladze, 1994b). We also mention the recent discovery of commensurate scale relations by Brodsky and Lu (1994, 1995). These relations allow one to connect several physical observables, providing important tests of QCD without scheme-scale ambiguity. \vspace{1cm} \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf Perturbative QCD series: How many loops should be evaluated?} The R-ratio in electron-positron annihilation is given within perturbation theory in the following form \begin{equation} R(s)=r_0\left(1+r_1(\frac{s}{\mu^2})\frac{\alpha_s(\mu)}{\pi} +r_2(\frac{s}{\mu^2}) (\frac{\alpha_s(\mu)}{\pi})^2 +r_3(\frac{s}{\mu^2}) (\frac{\alpha_s(\mu)}{\pi})^3 + ...\right). \label{eq:Rgeneral} \end{equation} Our further discussion is quite general and can be applied to other observables like R$_{\tau}$ or Higgs decay rates. We consider high enough energies, where R is a function of a single variable - the center-of-mass energy squared. Our aim is to evaluate pure QCD effects in R, which start with the term $O(\alpha_s)$, within the minimal subtraction prescription ('t Hooft, 1973). We should stress here that the calculational methods allowing one to evaluate perturbative corrections up to the four-loop order (up to the five-loop in some cases) is essentially based on some of the unique features of the MS prescription and our choice seems to be well justified. There is an ambiguity in the choice of renormalization scale parameter $\mu$. Usually we set $\mu^2=s$ and absorb the large logarithms in the definition of the running coupling. On the other hand, the choice $\mu^2=\chi s$ ($\chi \equiv e^{t}$) for all $\chi$ gives equivalent expansions. Evidently, the sum of ``all'' terms in eq.\ (\ref{eq:Rgeneral}) does not depend on the choice of $\mu$. However, in practice, we deal with truncated series, where the sum has a nontrivial dependence on the choice of renormalization parameter. Here we keep the ``natural'' choice $\mu^2=s$ and the ambiguity is transferred to the prescription $ \int d^4p \longrightarrow \int d^{4-2\varepsilon}p (\mu^2 e^{(t+O(\varepsilon))})^{\varepsilon}$. By changing $t$ one gets different MS type schemes. One can always reexpand (\ref{eq:Rgeneral}) in a new scheme (with a new $\Lambda$ in (\ref{eq:Asparametr}) ) and so redistribute the values of $r_{i}$ ($i > 1$). All these schemes are equivalent. On the other hand, a new scheme may be ``better'', but one can conclude this only based on the knowledge of remainders. The problem of scheme-scale ambiguity which, in fact, is a problem of remainders can be formulated as follows. {\em How does one choose (``optimize'') the scheme (or $\Lambda$) in order to make the remainder minimal in the series of the type (\ref{eq:Rgeneral}) for the given range of energy and what is the numerical uncertainty of the approximation (\ref{eq:Rgeneral})?} Here one should also distinguish the following two questions. First, what is the best accuracy to which the given quantity is calculable via perturbation theory? Second, what is the accuracy of the given approximation? A few notes are in order. It is known that perturbative QCD series are asymptotic ones. No reliable estimates of the remainders are known at present. However, it is known from the theory of asymptotic series (see, e.g., Dingle, 1973) that \begin{equation} \mid \sum_{i=1}^{\cal N}r_i\alpha^i(s)-R(s)\mid=R_{\cal N}\rightarrow \Delta R_{\mbox{\scriptsize{min}}}, \mbox{\hspace{2mm} when \hspace{2mm}} {\cal N}\rightarrow {\cal N}_{\mbox{\scriptsize{opt}}}. \label{eq:asestimate} \end{equation} This means that, the remainder $R_{\cal N}$ goes to its minimal value $\Delta R_{\mbox{\scriptsize{min}}}$ when the number of orders goes to its optimal value ${\cal N}_{\mbox{\scriptsize{opt}}}$. Inclusion of the next to ${\cal N}_{\mbox{\scriptsize{opt}}}$ orders will lead away from the correct value. It is known (see, e.g., Dingle, 1973) that for a sign-alternating asymptotic series the remainder can be estimated by the first neglected term (or by the last included term). However, it is still unknown if the QCD perturbative series has this character. We assume as a hypothesis that within QCD one can estimate the remainder by the first neglected or last included term. Now, the minimal possible error, which defines the best accuracy of the perturbation theory for the given quantity has an order of $\Delta R_{\mbox{\scriptsize{min}}}\sim r_{{\cal N}+1}\alpha^{{\cal N}+1}(s)$, ${\cal N}\rightarrow {\cal N}_{\mbox{\scriptsize{opt}}}$. Note that, both the number ${\cal N}_{\mbox{\scriptsize{opt}}}$ and the value of the $\Delta R_{\mbox{\scriptsize{min}}}$ depend on the range of energy for the given process. We once again emphasize that {\em the remainder depends on the choice of particular scheme and scale parameters and its estimate makes sense only for the ``optimized'' renormalization scheme which is unique for the given physical observable}. In fact, it was argued (Stevenson, 1984, 1994) that, the ``optimized'' series can still converge even when the series in any fixed renormalization scheme is factorially divergent, if the ``optimized'' couplant shrinks in higher orders (see also Buckley, Duncan and Jones, 1993). However, whether this applies to QCD is unknown. \vspace{1cm} \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf $R(s)$ within the one parametric family of the MS type schemes and scale ambiguity problem} Using the results of our four-loop calculations, we obtain the analytical result for $R(s)$ with perturbative coefficients explicitly depending on the scheme-scale parameter (Surguladze and Samuel, 1993) \begin{eqnarray} \lefteqn{\hspace{-7mm}R(s,t)=R_0+\frac{\alpha_s(s,t)}{\pi}R_1 +\biggl(\frac{\alpha_s(s,t)}{\pi}\biggr)^2 (R_2+\beta_0 R_1 t) } \nonumber\\ && \hspace{33mm} +\biggl(\frac{\alpha_s(s,t)}{\pi}\biggr)^3 [R_3-\frac{\pi^2}{3}\beta_0^2R_1 +(2\beta_0 R_2 + \beta_1 R_1)t+\beta_0^2 R_1 t^2]. \nonumber\\ \label{eq:Rstgen} \end{eqnarray} Recalling the values of the $\overline{\mbox{MS}}$ perturbative coefficients $R_i$ from eqs.\ (\ref{eq:Rexpans}) and (\ref{eq:Ranalytic0}) and the $\beta_i$ coefficients from eq.\ (\ref{eq:beta}), we obtain numerically \begin{eqnarray} \lefteqn{\hspace{-14mm}R(s,t)= 3\sum_{f}Q_{f}^2\biggl\{ 1+ \frac{\alpha_s(s,t)}{\pi} +\biggl(\frac{\alpha_s(s,t)}{\pi}\biggr)^2 [(1.9857+2.75t)-N(0.1153+0.1667t)]} \nonumber\\ && \quad \hspace{3cm} +\biggl(\frac{\alpha_s(s,t)}{\pi}\biggr)^3 [(-6.6369+17.2964t+7.5625t^2) \nonumber\\ && \quad \hspace{50mm} -N(1.2001+2.0877t+0.9167t^2) \nonumber\\ && \quad \hspace{50mm} +N^{2}(-0.0052+0.0384t+0.0278t^2)]\biggr\} \nonumber\\ && \quad \hspace{12mm} -\biggl(\sum_{f}Q_f\biggr)^2 \biggl(\frac{\alpha_s(s,t)}{\pi}\biggr)^3 1.2395 +O(\alpha_s^4), \label{eq:Rstnumer} \end{eqnarray} where $\alpha_{s}(s,t)$ can be parametrized in the form of (\ref{eq:Asparametr}) with $\mu=s$ and $\Lambda\rightarrow \Lambda_t=e^{-t/2}\Lambda_{\overline{\mbox{\scriptsize{MS}}}}$. Obviously, $t=0$ corresponds to the $\overline{\mbox{MS}}$ scheme ( eq.\ (\ref{eq:Rnumerical}) ). $t=\ln 4\pi-\gamma$ will transform the result to the original MS scheme ('t Hooft, 1973). ( eq.\ (\ref{eq:RMSnumerical}) ). t=-2 corresponds to the G scheme introduced by Chetyrkin and Tkachov (1979, 1981) ( eq.\ (\ref{eq:RGnumerical}) ). Note that because of a one-parametric nature of the MS prescription, the $t$-dependent terms in eq.\ (\ref{eq:Rstnumer}) would represent also the scale dependence of the perturbative coefficients within the $\overline{\mbox{MS}}$ if one changes $t\rightarrow \log\mu^2/s$ and takes $\alpha_s(s,t)$ with $s$ replaced by $\mu^2$ and $t=0$. Several approaches were suggested to deal with the scheme-scale-remainder problem. Among them we consider the following ones. {\it Fastest Apparent Convergence} (FAC) (Grunberg, 1980, 1982, 1984), where the next to leading perturbative correction is absorbed in the definition of the ``effective'' running coupling and the scheme-scale parameter is fixed accordingly. {\it Principle of Minimal Sensitivity} (PMS) of the approximant to the variation of nonphysical parameters (Stevenson, 1981a,b, 1982, 1984; see also Mattingly and Stevenson, 1992, 1994). {\it Brodsky-Lepage-Mackenzie} (BLM) approach (Brodsky, Lepage and Mackenzie, 1983), which suggests one fix the scale by the size of the quark vacuum polarization effects resulting in the independence of the next to leading order perturbative correction of the number of quark flavors $N$. For discussions of the above scheme-scale setting methods see Celmaster and Stevenson (1983), Brodsky and Lu (1992), and Stevenson (1992). The optimization of perturbation theory has previously been studied by Kramer and Lampe (1988), and Bethke (1989) for jet cross sections in electron positron annihilation. The optimized perturbation theory is tested for different physical quantities in QED and QCD by Field (1993). The scale ambiguity problem has been considered by Lu and de Melo (1991) for the $\phi^{3}$ model. The scheme-scale ambiguity problem for the quantities $R(s)$ and $R_{\tau}$ has been discussed by Maxwell and Nicholls (1990), Chyla, Kataev and Larin (1991), and Grunberg and Kataev (1992). Further study of the PMS method has been done in Raczka (1995). We apply the above methods to eq.\ (\ref{eq:Rgeneral}) and we find a scale which gives good results for all criteria considered (Surguladze and Samuel, 1993). We start by noting that, in general, the renormalizatin scheme-scale dependence of perturbative results are parametrized by the scale parameter, say, $\mu$ and the renormalization prescription dependent coefficients of $\beta$ function (Stevenson, 1981a,b). We should stress however, that the $\beta$ function is independent of any modification of the MS type prescriptions, but starting from $\beta_2$, the coefficients of $\beta$ function do depend on the particular choice of subtraction prescription other than MS. In order to better visualize our discussion, we consider first the optimization procedures within the MS prescription. In other words, we fix the scheme dependent perturbative coefficients of $\beta$ function to their MS values and consider only the scale variation. In Fig.\ 12 we have plotted $r_{3}(t)$ for different $N$ (see eqs.\ (\ref{eq:Rgeneral}) and (\ref{eq:Rstnumer})). As one can see, within the region $t\sim(-1.5,-0.5)$ $r_3$ has a very weak dependence on the number of flavors $N$ as well as on the parameter $t$. \vspace{9cm} \begin{center} FIG.\ 12. $r_{3}(t)$ for different $N$ \end{center} \vspace{2mm} Corresponding to the three-loop coefficient $r_2(t)$, straight lines intersect in one point for $t\approx -0.7 $, which is obvious from eq.\ (\ref{eq:Rstnumer}). This value corresponds to the BLM result (Brodsky, Lepage and Mackenzie, 1983) $\mu^2=\mu^{2}_{\overline{\mbox{\scriptsize MS}}}e^{0.710}$, and at this scale the flavor dependence is absorbed into the definition of the coupling. In Fig.\ 13 we have plotted the dependence of the partial sums \begin{displaymath} R_{n}(t)=\sum_{m=1}^{n}r_{m}(t)(\alpha_s/\pi)^m, \hspace{5mm} n=1,2,3 \end{displaymath} on the parameter $t$. Here the parametrisation (\ref{eq:Asparametr}) was used, $\log s/\Lambda_{\overline{\mbox{\scriptsize MS}}}^2=9$ and $N=5$. The general picture does not change for other reasonable values of $\log$ and $N$. One can see that PMS (Stevenson, 1981) works perfectly for a wide range of the logarithmic scale parameter $t\sim(-1,+3)$ for the four-loop approximant and $t\sim(-2,0)$ for the three-loop approximant. A similar analysis at the three-loop level was done by Radyushkin (1983). According to the above analysis we found that the BLM scale $t=-0.710$ is good at the four-loop level as well (Fig.\ 12) and this value is within minimal sensitivity region (Fig.~13). \vspace{77mm} \begin{center} FIG.\ 13. The approximants $R_{n}$ vs the scale parameter $t$ \end{center} \vspace{2mm} \noindent Moreover, we found that if the $t$-parameter is chosen in the following analytical form $t=4\zeta(3)-11/2+O(\varepsilon)$, which is equivalent to the definition of a new, say, $\widetilde{\mbox{MS}}$ modification of the MS scheme \begin{equation} \Lambda_{\widetilde{\mbox{\scriptsize MS}}}=\mbox{exp} [-2\zeta(3)+11/4+O(\varepsilon)] \Lambda_{\overline{\mbox{\scriptsize MS}}}, \label{eq:mstilda} \end{equation} then the $N$ dependence and the $\zeta(3)$ terms cancel exactly at the 3-loop level. As a result, $r_2=1/12$. Within this scheme the four-loop correction is almost independent of the number of flavors. The full result for the R-ratio for the {\em arbitrary} number of flavors can be written in the following simple form \begin{equation} R(s)=3\sum_{f}Q_{f}^2\biggl[ 1+\frac{\alpha_s}{\pi} +\frac{1}{12}\biggl(\frac{\alpha_s}{\pi}\biggr)^2 -\biggl(\frac{\alpha_s}{\pi}\biggr)^3 (16.2 \pm 0.5)\biggr] -\biggl(\sum_{f}Q_f\biggr)^2 \biggl(\frac{\alpha_s}{\pi}\biggr)^3 1.2 +O(\alpha_s^4) \label{eq:Rsmstilda} \end{equation} where the small uncertainty $\pm 0.5$ stands for the remainder dependence on the number of flavors at $O(\alpha_s^3)$ for all physically reasonable $N$ and is completely negligible for phenomenology. The last term is also very small $\sim 0.4(\alpha_{s}/\pi)^3$. The running coupling can be parametrized in the standard form (\ref{eq:Asparametr}) with $\Lambda_{\widetilde{\mbox{\scriptsize MS}}} =1.41\Lambda_{\overline{\mbox{\scriptsize MS}}}$. Using the FAC approach (Grunberg, 1980, 1982, 1984), we rewrite eq.\ (\ref{eq:Rsmstilda}) as follows. \begin{equation} R(s)= 3\sum_{f}Q_{f}^2\left[ 1+\frac{\alpha_{s}^{\mbox{\scriptsize eff}}}{\pi} +O(\alpha_s^3)\right], \label{eq:Rseffect} \end{equation} where the 3-loop correction is absorbed into the definition of the effective coupling given by eq.\ (\ref{eq:Asparametr}) with the $\Lambda$ replaced by \begin{displaymath} \Lambda_{\mbox{\scriptsize eff}} \approx \Lambda_{\widetilde{\mbox{\scriptsize MS}}} \mbox{exp}\left(\frac{1}{2\beta_0} \frac{r_2}{r_1}\right) \approx 1.02\Lambda_{\widetilde{\mbox{\scriptsize MS}}}. \end{displaymath} As one can see, the new scheme $\widetilde{\mbox{MS}}$ almost coincides with the effective one and the fastest convergence is guaranteed within the wide range of energy defined by the renormalization group invariant criteria \begin{displaymath} \frac{s}{\Lambda_{\mbox{\scriptsize eff}}^{2}} \sim \frac{s}{\Lambda_{\widetilde{\mbox{\scriptsize MS}}}^{2}} \gg 1. \end{displaymath} The similar analyses can be done for the semi-hadronic decay rates of the $\tau$ lepton calculated to $O(\alpha_s^3)$ in section 7. The result for the ratio $R_{\tau}$ in the $\widetilde{\mbox{MS}}$ scheme reads \begin{equation} R^{\tau}= 3(0.998\pm 0.002)\left[ 1+\frac{\alpha_s(M_{\tau}^2)}{\pi} +3.65\left(\frac{\alpha_s(M_{\tau}^2)}{\pi}\right)^2 +9.83\left(\frac{\alpha_s(M_{\tau}^2)}{\pi}\right)^3 \right] +O(\alpha_s^4) \label{eq:Rtaumstilde} \end{equation} and to be compared to eq.\ (\ref{eq:Rtaunumer}). Note that the $\alpha_s(M_Z)$ is parametrized with the $\Lambda_{\widetilde{\mbox{\scriptsize MS}}} =1.41\Lambda_{\overline{\mbox{\scriptsize MS}}}$. In Fig.\ 14 we plot one-, two- and three-loop approximants to the $\Gamma_{H\rightarrow b\overline{b}}$ in terms of the running quark mass (eqs.\ (\ref{eq:Danalyt})-(\ref{eq:GHtot}), with $N=5$ and $m_f=m_b$) vs. the scale parameter $t$ (Surguladze, 1994b). \newpage \mbox{} \vspace{75mm} \begin{center} FIG.\ 14. The approximants of the $\Gamma_{H\rightarrow b\overline{b}}$ vs the scale parameter $t$ \end{center} \vspace{2mm} One can see that the higher order corrections diminish the scale dependence from 40\% to nearly 5\%. The solid curve, corresponding to the three-loop result, became flat in the wide range of the logarithmic scale parameter $t$. Moreover, the choice $t=0$ ($\overline{\mbox{MS}}$-scheme) satisfies Stevenson's {\it Principle of Minimal Sensitivity} (Stevenson, 1981). Let us now try to estimate the theoretical uncertainty in calculations of $R$ by the last included term in the corresponding perturbative expansion. We get for the QCD contribution within the $\widetilde{\mbox{MS}}$ scheme the following result. \begin{equation} \delta^{\widetilde{\mbox{\scriptsize MS}}}_{\mbox{\scriptsize QCD}} \equiv \frac{R(s)-r_0}{r_0}=\frac{\alpha_s}{\pi} +\frac{1}{12}\left(\frac{\alpha_s}{\pi}\right)^2 -(16.2 \pm 0.5) \left(\frac{\alpha_{s}}{\pi}\right)^3 \pm (\delta_{\mbox{\scriptsize QCD}}^{\mbox{\scriptsize err}}=4\%). \label{eq:Rserrorestincl} \end{equation} The analysis of Fig.\ 13 shows that the deviation of the four-loop approximant from the constant is also about $4\%$ within a reasonably wide range of the t-parameter. This is consistent with Stevenson's principle. One should note that the above error estimate is only for the massless quark limit. There are several different types of additional contributions, including those due to nonvanishing quark masses. This may change the above error estimate. All of the necessary information on the status of the additional corrections can be found in Kniehl (1994b, 1995b). As we have already mentioned, recently Brodsky and Lu (1994, 1995) have found the relations between the effective couplings $\alpha_{\mbox{\tiny A}}$ and $\alpha_{\mbox{\tiny B}}$ for the physical observables A and B in the following form. \begin{equation} \alpha_{\mbox{\tiny A}}(\mu_{\mbox{\tiny A}}) =\alpha_{\mbox{\tiny B}}(\mu_{\mbox{\tiny B}}) \biggl(1+r_{\mbox{\tiny A/B}}\frac{\alpha_{\mbox{\tiny B}}}{\pi} +\cdots\biggr). \label{eq:CSR} \end{equation} The ratio of the scales of the corresponding processes $\mu_{\mbox{\tiny A}}/\mu_{\mbox{\tiny B}}$ is chosen according to the BLM scale setting prescription so that $r_{\mbox{\tiny A/B}}$ is independent of the number of flavors. Thus, evolving $\alpha_{\mbox{\tiny A}}$ and $\alpha_{\mbox{\tiny B}}$, they pass the quark thresholds at the same scale. It is shown that the relative scales satisfy the transitivity rule \begin{displaymath} \frac{\mu_{\mbox{\tiny A}}}{\mu_{\mbox{\tiny B}}} =\frac{\mu_{\mbox{\tiny A}}}{\mu_{\mbox{\tiny C}}} \times\frac{\mu_{\mbox{\tiny C}}}{\mu_{\mbox{\tiny B}}}. \end{displaymath} So, C may correspond to any intermediate theoretical scheme such as MS, $\overline{\mbox{MS}}$, etc. and the perturbative results can be tested without a reference to them. One of the impressive results of this method is a surprisingly simple relation between the effective couplings for the quantities $R$ and $R_{\tau}$ to the next-to-next leading order (Brodsky and Lu, 1994, 1995) \begin{displaymath} \frac{\alpha_{\tau}(M_{\tau})}{\pi}= \frac{\alpha_{R}(\mu)}{\pi}, \hspace{9mm} \mu=M_{\tau}\mbox{exp}\biggl[-\frac{19}{24}-\frac{169}{128} \frac{\alpha_R(M_{\tau})}{\pi}\biggr]. \end{displaymath} For more details and the relations between various other observables we refer to the original works by Brodsky and Lu (1994, 1995). \vspace{1cm} \renewcommand{\thesection}{\arabic{section}} \subsection{\tenbf On scheme invariant analyses} Let us now outline the original method of scheme-invariant analyses for the perturbation theory results by Stevenson (1981a,b, 1982, 1984). We note first, that our analyses of perturbation series for $R(s)$ and $R_{\tau}$ has been done in the previous subsection within the one parametric family of the MS type schemes, where all $\beta$ function coefficients are the same for any modification of MS. In the PMS method, renormalizarion scale and scheme dependence is parametrized by the scale parameter $\mu/\Lambda$ and the scheme dependent coefficients of the $\beta$ function $\beta_2$, $\beta_3,\cdots$. Then the {\it Principle of Minimal Sensitivity} is applied to the variation of the above parameters and to $O(\alpha_s^3)$ the ``optimized'' scheme corresponds to a flat two dimensional surface. Our curve for $R_3$ in Fig.\ 13 is just a one-dimensional slice at the particular MS value of the $\beta_2$. The main points of the PMS formalism is as follows. (For the scheme invariant analyses of $R(s)$ to $O(\alpha_s^3)$ see Mattingly and Stevenson, 1994). To use familiar standard notation, we rewrite eq.\ (\ref{eq:RGfunctions}) for the couplant $a\equiv\alpha_s(\mu)/\pi$ \begin{equation} b\frac{\partial a}{\partial \tau} =-b a^2(1+ca+c_2a^2+\cdots), \label{eq:ab} \end{equation} where \begin{equation} \tau=b\ln\frac{\mu}{\Lambda}, \hspace{4mm} b=2\beta_0, \hspace{4mm} c=\frac{\beta_1}{\beta_0} \label{eq:ta} \end{equation} and for any modification of the minimal subtraction prescription, the scheme dependent coefficient $c_2=\beta_2/\beta_0$. The scheme and scale can now be parametrized by the quantities $RS\equiv (\tau,c_2,c_3,...)$. The {\it Principle of Minimal Sensitivity} can be written as \begin{equation} \frac{dR_n}{d(\tau;c_2,c_3,...)}=0. \label{eq:PMS} \end{equation} The number of scheme-scale parameters in the above equation is strongly correlated with $n$. Indeed, it is not difficult to show that the following self-consistency condition should hold for the $n$th approximant \begin{equation} \frac{\partial R_n}{\partial(RS)}=O(a^{n+1}). \label{eq:SCC} \end{equation} This shows that the perturbative coefficients $r_i$ can depend on renormalization scheme only through parameters $\tau;c_2,...,c_{i-1}$. Applying the {\it Principle of Minimal Sensitivity} in a form (\ref{eq:PMS}) to the approximants $R_2$ and $R_3$ and taking into account (\ref{eq:SCC}), one finds that the quantities \begin{displaymath} \rho_1\equiv \tau-r_2, \end{displaymath} \begin{equation} \rho_2\equiv r_3+c_2-\biggl(r_2+\frac{c}{2}\biggr)^2 \label{eq:RSinv} \end{equation} are renormalization scheme independent. Similar invariants can be constructed at each order of perturbation theory. The choice of $\tau$ as a function of the ratio $\mu/\Lambda$ emphasizes that the renormalization scheme dependence involves only the ratio of these quantities and the optimization deals with $\tau$ but not $\mu$. The ``optimal'' values of renormalization scheme parameters $\overline{\tau}$ and $\overline{c_2}$ are defined by the following equations. To $O(\alpha_s^2)$, \begin{equation} \frac{d R_2(\tau)}{d\tau}\biggr|_{\tau=\overline{\tau} }=0. \label{eq:tauopt} \end{equation} To $O(\alpha_s^3)$, \begin{equation} \frac{\partial R_3(\tau,c_2)}{\partial\tau}\biggr|_{\tau=\overline{\tau}}=0, \end{equation} \begin{equation} \frac{\partial R_3(\tau,c_2)}{\partial c_2}\biggr|_{c_2=\overline{c}_2}=0. \label{eq:taucopt} \end{equation} Solving the above equations along with eqs.\ (\ref{eq:RSinv}) for the renormalization scheme invariants and eq.\ (\ref{eq:ab}) for the couplant with the truncated MS $\beta$ function, using the $\overline{\mbox{MS}}$ values of $r_2$ and $r_3$, one finds the ``optimized'' values of $\overline{\tau}$, $\overline{c}_2$ and corresponding ``optimized'' approximants to $O(\alpha_s^3)$. The theoretical error can be estimated, as in the previous subsection, by the last calculated term. One obtains the following ``optimized'' result for the QCD contribution in $R(34\mbox{ GeV})$ in the massless quark limit (Mattingly and Stevenson, 1994; Stevenson, 1994). \begin{equation} \delta_{\mbox{\tiny QCD}}^{\mbox{\tiny PMS}} = 0.051 \pm 0.001. \label{eq:PMSres} \end{equation} It is important to note that the above optimization procedure yields a negative value for the $\rho_2$ invariant. This results in the existence of a solution of equation \begin{equation} \frac{7}{4}+c\overline{a}^{\ast}+3\rho_2(\overline{a}^{\ast})^2=0 \label{eq:IRfix} \end{equation} with respect to $\overline{a}^{\ast}$ - the value of the couplant for which the optimized third order $\beta$ function vanishes. This allows, in principle, to do some analyses for $R(s)$ at the low energies $\sqrt{s}\rightarrow 0$ (Mattingly and Stevenson, 1992). Finally, we also mention that the FAC approach (Grunberg, 1980, 1982, 1984) is a special case of the PMS (Stevenson, 1981a,b, 1982, 1984) method. Indeed, in the FAC approach all higher order approximants are equal to the effective couplant (compare to eqs.\ (\ref{eq:tauopt}) and (\ref{eq:taucopt}) ). {}From eqs.\ (\ref{eq:RSinv}) one gets $\rho_1=\tau$ and $\rho_2=c_2$ in the FAC approach. \vspace{7mm} \section*{\bf Conclusions} \addtocontents{toc}{{\bf Conclusions} \hspace{124mm} {\bf 72}} \indent In the present article we reviewed the current development of calculational methods, algorithms and computer programs which allow one to evaluate the characteristics of the phenomenologically important physical processes to higher orders of perturbative QCD. We have considered $Z\rightarrow \mbox{hadrons}$, $\tau^{-} \rightarrow \nu_{\tau} +\mbox{hadrons}$, $H \rightarrow \mbox{hadrons}$. The described methods are applicable to a wide class of calculational problems of modern high energy physics. We outlined the analytical three- and four-loop calculations for the above mentioned processes. The methods of analytical perturbative calculations available at present allow, in principle, one to evaluate various decay rates, cross-sections, coefficient functions in the operator product expansion, renormalization group functions etc. up to and including five-loop level. This would correspond, for instance, the decay rate in the process $Z \rightarrow \mbox{hadrons}$ to $O(\alpha_s^4)$. It seems that such a high order will completely fit the experimental state of the problem in the observable future. Indeed, for example, the 4\% estimate of the theoretical error for the decay rate of $Z$-boson is based on the $O(\alpha_s^3)$ calculation. The present experimental error at LEP is about 5\%. The involvement of the heavier quarks in the physical processes makes it necessary to develop methods for calculation of the Feynman graphs with the propagators of massive particles. The expansion in terms of large or small masses may not always give satisfactory results. The problem of the renormalization group ambiguity of the perturbation theory results and various methods for resummation of higher order corrections is a subject of growing interest and discussions in the literature. The future development of analytical programming tools towards the full automation of high order calculations would be welcome. This would greatly reduce the chance of errors in the calculations. On the other hand, the computer package with full implementation of the algorithm of high order analytical perturbative calculations would make it realistic to step up by one more order. We recognize that it is unavoidable that some of the relevant references have not been mentioned. We assure the reader that this is only due to our unintentional ignorance. \vspace{1cm} \noindent {\bf ACKNOWLEDGMENTS} \indent It is a pleasure to thank D.\ Soper for numerous discussions and his support. We are grateful to N.\ Deshpande and R.\ Hwa for discussions on the present status of the Standard Model. We thank E.\ Braaten, S.\ Brodsky and B.\ Kniehl for their comments at various stages of this work. We would especially like to thank P.\ Stevenson for reading the manuscript, illuminating discussions, suggestions and correcting the errors. L.R.S. would like to thank members of the experimental high energy physics laboratory at the University of Oregon, especially J.\ Brau, R.\ Fray and D.\ Strom for encouraging discussions. L.R.S. would like to thank N.\ S.\ Amaglobeli, V.\ A.\ Matveev, V.\ A.\ Rubakov and A.~N.~Tavkhelidze for their interest in our work and their support, the members of the Theory Division of the Moscow Institute for Nuclear Research for collaboration on various problems which further became topics of the present article, and the members of the Department of High Energy Physics, Tbilisi State University for discussions. We are grateful to C.\ Quigg for encouraging us to write this review. \vspace{3mm} \noindent This work was supported by the U.S. Department of Energy under grant No. DE-FG06-85ER-40224 and under grant No. DE-FG05-84ER40215. \newpage \noindent {\bf REFERENCES} \vspace{6mm} \noindent Abers, E.\ S., and B.\ W.\ Lee, 1973, Phys.\ Rep.\ {\bf 9,} 1.\\ Abbot, L., 1980, Phys.\ Rev.\ Lett.\ {\bf 44,} 1569.\\ Acharya, A., and B.\ P.\ Nigam, 1978, Nucl.\ Phys.\ {\bf B 141,} 178.\\ Acharya, A., and B.\ P.\ Nigam, 1985, Nuovo Cim.\ {\bf A 88,} 293.\\ Adler, S.\ L., 1972, Phys.\ Rev.\ {\bf D 5,} 3021.\\ Adler, S.\ L., 1974, Phys.\ Rev.\ {\bf D 10,} 3714.\\ Altarelli, G., 1982, Phys.\ Rep.\ {\bf 81,} 1.\\ Altarelli, G., 1989, Annu.\ Rev.\ Nucl.\ Sci.\ {\bf 39,} 357.\\ Altarelli, G., 1992, ``QCD and experiment: status of $\alpha_s$,'' CERN preprint No.\ TH.6623/92.\\ Altarelli, G., P.\ Nason, and G.\ Ridolfi, 1994, A study of ultraviolet renormalon ambiguities in the determination of $\alpha_s$ from $\tau$ decay,'' CERN preprint No.\ TH.7537/94.\\ Appelquist, T., and J.\ Carazzone, 1975, Phys.\ Rev.\ {\bf D 11,} 2856.\\ Appelquist, T., and D.\ Politzer, 1975, Phys.\ Rev.\ Lett.\ {\bf 34,} 43.\\ Ashmore, J.\ F., 1972, Nuov.\ Cimm.\ Lett.\ {\bf 4,} 289.\\ Baker, M., and K.\ Johnson, 1969, Phys.\ Rev.\ {\bf 183,} 1292.\\ Baker, M., and K.\ Johnson, 1971, Phys.\ Rev.\ {\bf D 3,} 2541.\\ Banyai, L., S.\ Marculescu, and T.\ Vescan, 1974, Lett.\ Nuov.\ Cim.\ {\bf 11,} 151.\\ Barclay, D.\ T., and C.\ J.\ Maxwell, 1992a, Phys.\ Rev.\ Lett.\ {\bf 69,} 3417.\\ Barclay, D.\ T., and C.\ J.\ Maxwell, 1992b, Phys.\ Rev.\ {\bf D 45,} 1760.\\ Bardeen, W., A.\ Buras, D.\ Duke, and T.\ Muta, 1978, Phys.\ Rev.\ {\bf 18,} 3998.\\ Barger, V.\ D., and R.\ J.\ N.\ Phillips, 1987, {\it Collider Physics,} Frontiers in Physics Series {\bf 71} (Addison-Wesley).\\ Barnett, M.\ R., M.\ Dine, and L.\ McLerran, 1980, Phys.\ Rev.\ {\bf D 22,} 594.\\ Barnett, M.\ R., H.\ E.\ Haber, and D.\ E.\ Soper, 1988, Nucl.\ Phys.\ {\bf B 306,} 697.\\ Bechi, C., S.\ Narison, E.\ de Rafael, and F.\ Yndurain, 1981, Z.\ Phys.\ {\bf 1981,} 335.\\ Bernreuter, W., and W.\ Wetzel, 1982, Nucl.\ Phys.\ {\bf B 197,} 228.\\ Bethke, S., 1989, Z.\ Phys.\ {\bf C 43,} 331.\\ Bethke, S., 1992, in {\sl Proceedings of the 26th International Conference on High Energy Physics} (Dallas, USA), p.\ 81.\\ Bethke, S., and J.\ E.\ Pilcher, 1992, Annu.\ Rev.\ Nucl.\ Sci. {\bf 42,} 251.\\ Bjorken, J.\ D., 1968, in {sl Proceedings of 1967 Int. School of Physics, Enrico Fermi, Course 41,} Varenna, Italy (Academic Press, New York), p.\ 55.\\ Bjorken, J.\ D., 1969, Phys.\ Rev.\ {\bf 179,} 1547.\\ Bloch, F., and A.\ Nordsieck, 1937, Phys.\ Rev.\ {\bf 52,} 54.\\ Bogolyubov, N.\ N., and O.\ S.\ Parasyuk, 1955a, Dokl. Akad. Nauk SSSR [Sov.\ Phys.\ Dokl.] {\bf 100,} 25.\\ Bogolyubov, N.\ N., and O.\ S.\ Parasyuk, 1955b, Dokl.\ Akad.\ Nauk SSSR [Sov.\ Phys.\ Dokl.] {\bf 100,} 429.\\ Bogolyubov, N.\ N., and O.\ S.\ Parasyuk, 1956, Izv.\ Akad.\ Nauk SSSR, ser.\ matem.\ {\bf 20,} 585.\\ Bogolyubov, N.\ N., and O.\ S.\ Parasyuk, 1957, Acta Mathem.\ {\bf 97,} 227.\\ Bogolyubov, N.\ N., and D.\ V.\ Shirkov, 1955, Dokl.\ Akad.\ Nauk SSSR [Sov.\ Phys.\ Dokl.] {\bf 103,} 203.\\ Bogolyubov, N.\ N., and D.\ V.\ Shirkov, 1956a, Sov.\ Phys.-JETP (translation of Zh.\ Eksp.\ Teor.\ Fiz.) {\bf 30,} 77.\\ Bogolyubov, N.\ N., and D.\ V.\ Shirkov, 1956b, Nuov.\ Cim. {\bf 3,} 845.\\ Bogolyubov, N.\ N., and D.\ V.\ Shirkov, 1980, {\sl Introduction to the Theory of Quantized Fields} (John Wiley \& Sons, Inc.).\\ Bogolyubov, N.\ N., B.\ V.\ Struminsky, and A.\ N.\ Tavkhelidze, 1965, JINR report No.\ JINR-D-1968.\\ Bollini, C.\ G., and J.\ J.\ Giambiagi, 1972, Phys.\ Lett.\ {\bf 40 B,} 566.\\ Bonneau, G., 1980, Phys.\ Lett.\ {\bf 96 B,} 147.\\ Boos, E.\ E., and A.\ I.\ Davydychev, 1992, Theor.\ Math.\ Phys.\ {\bf 89,} 1052.\\ Braaten, E., 1988, Phys.\ Rev.\ Lett.\ {\bf 60,} 1606.\\ Braaten, E., and J.\ P.\ Leveille, 1980, Phys.\ Rev.\ {\bf D 22,} 715.\\ Braaten, E., C.\ S.\ Li, 1990, Phys.\ Rev.\ {\bf D 42,} 3888.\\ Braaten, E., S.\ Narison, and A.\ Pich, 1992, Nucl.\ Phys.\ {\bf B 373,} 581.\\ Broadhurst, D.\ J., and S.\ G.\ Generalis, 1982, ``Pseudoscalar QCD sum rules,'' Open University preprint No.\ {\bf OUT-4102-8}.\\ Broadhurst, D.\ J., and S.\ G.\ Generalis, 1985, Phys.\ Lett.\ {\bf 165 B,} 175.\\ Broadhurst, D.\ J., N.\ Gray and K.\ Schilcher, 1991, Z.\ Phys.\ {\bf C 52,} 111.\\ Broadhurst, D.\ J., A.\ L.\ Kataev, and O.\ V.\ Tarasov, 1993, Phys.\ Lett.\ {\bf B 298,} 445.\\ Broadhurst, D.\ J., {\it et al.}, 1994, Phys.\ Lett.\ {\bf B 329,} 103.\\ Brock, R., {\it et al.}, CTEQ Collaboration, 1993, {\sl Handbook of Perturbative QCD,} Edited by G.\ Sterman.\\ Brodsky, S.\ J., 1993, ``New perspective in Quantum Chromodynamics,'' SLAC preprint No.\ SLAC-PUB-6304.\\ Brodsky, S.\ J., and H.\ J.\ Lu, 1992, ``On the selfconsistency of scale setting methods,'' SLAC preprint No.\ SLAC-PUB-6000.\\ Brodsky, S.\ J., and H.\ J.\ Lu, 1994, ``Commensurate scale relations: precise tests of Quantum Chromodynamics without scale or scheme ambiguity,'' SLAC preprint No.\ SLAC-PUB-6683.\\ Brodsky, S.\ J., and H.\ J.\ Lu, 1995, Phys.\ Rev.\ {\bf D 51,} 3652.\\ Brodsky, S.\ J., and G.\ R.\ Ferrar, 1973, Phys.\ Rev.\ Lett.\ {\bf 31,} 1153.\\ Brodsky, S.\ J., G.\ P.\ Lepage, and P.\ B.\ Mackenzie, 1983, Phys.\ Rev.\ {\bf D 28,} 228.\\ Brown, L.\ S., and L.\ G.\ Yaffe, 1992, Phys.\ Rev.\ {\bf D 45,} 398.\\ Brown, L.\ S., L.\ G.\ Yaffe, and C.\ X.\ Zhai, 1992, ``Large order perturbation theory for the electromagnetic current-current correlation function,'' Washington University preprint No.\ UW-PT-92-07.\\ Buckley, I.\ R.\ C., A.\ H.\ Duncan, and H.\ F.\ Jones, 1993, Phys.\ Rev.\ {\bf D 47,} 2554.\\ Buras, A.\ J., E.\ G.\ Floratos, D.\ A.\ Ross, and C.\ T.\ Sachrajda, 1977, Nucl.\ Phys.\ {\bf B 131,} 308.\\ Callan, C., 1970, Phys.\ Rev.\ {\bf D 2,} 1541.\\ Caswell, W.\ E., and F.\ Wilczek, 1974, Phys.\ Lett.\ {\bf B 49,} 291.\\ Celmaster, W., and R.\ G.\ Gonsalves, 1979, Phys.\ Rev.\ {\bf D 20,} 1420.\\ Celmaster, W., and R.\ G.\ Gonsalves, 1980, Phys.\ Rev.\ Lett.\ {\bf 44,} 560.\\ Celmaster, W., and P.\ M.\ Stevenson, 1983, Phys.\ Lett.\ {\bf B 125,} 493.\\ Chetyrkin, K.\ G., 1988, Teor.\ Mat.\ Fiz.\ {\bf 76,} 207 [Theor.\ Math.\ Phys.\ {\bf 76,} 809 (1988)].\\ Chetyrkin, K.\ G., 1991, ``Combinatorics of R, R$^{-1}$, and R$^{\ast}$ operations and asymptotic expansions of Feynman integrals in the limit of large momenta and masses,'' Max Planck Institute preprint No.\ MPI-PAE/PTh 13/91.\\ Chetyrkin, K.\ G., 1992, Phys.\ Lett.\ {\bf B 282,} 221.\\ Chetyrkin, K.\ G., 1993a, Phys.\ Lett.\ {\bf B 307,} 169.\\ Chetyrkin, K.\ G., 1993b, ``Possible and impossible in multiloop renormalization group,'' Karlsruhe University preprint No.\ TTP93-37.\\ Chetyrkin, K.\ G., S.\ G.\ Gorishny, and V.\ P.\ Spiridonov, 1985, Phys.\ Lett.\ {\bf B 160,} 149.\\ Chetyrkin, K.\ G., S.\ G.\ Gorishny ,and F.\ V.\ Tkachov, 1982, Phys.\ Lett.\ {\bf B 119,} 407.\\ Chetyrkin, K.\ G., A.\ L.\ Kataev, and F.\ V.\ Tkachov, 1979, Phys.\ Lett.\ {\bf 85,} 277.\\ Chetyrkin, K.\ G., A.\ L.\ Kataev, and F.\ V.\ Tkachov, 1980, Nucl.\ Phys.\ {\bf B 174,} 345.\\ Chetyrkin, K.\ G., and A.\ Kwiatkowski, 1993, Z.\ Phys.\ {\bf C 59,} 525.\\ Chetyrkin, K.\ G., and A.\ Kwiatkowski, 1995, ``Second order QCD corrections to scalar and pseudoscalar Higgs decays into massive bottom quarks,'' LBL preprint No.\ LBL-37269.\\ Chetyrkin, K.\ G., and J.\ H.\ K\"{u}hn, 1990, Phys.\ Lett.\ {\bf B 248,} 359.\\ Chetyrkin, K.\ G., and J.\ H.\ K\"{u}hn, 1992, Phys.\ Lett.\ {\bf B 282,} 359.\\ Chetyrkin, K.\ G., J.\ H.\ K\"{u}hn, and A.\ Kwiatkowski, 1992, Phys.\ Lett.\ {\bf B 282,} 221.\\ Chetyrkin, K.\ G., and F.\ V.\ Tkachov, 1979, ``New approach to evaluations of multiloop Feynman diagrams,'' Moscow Institute for Nuclear Research preprint No. P-0018.\\ Chetyrkin, K.\ G., and F.\ V.\ Tkachov, 1981, Nucl.\ Phys.\ {\bf B 192,} 159.\\ Chetyrkin, K.\ G., and F.\ V.\ Tkachov, 1982, Phys.\ Lett.\ {\bf B 114,} 340.\\ Chyla, J., A.\ L.\ Kataev, and S.\ A.\ Larin, 1991, Phys.\ Lett.\ {\bf B 267,} 269.\\ Cicuta, G.\ M., and E.\ Montaldi, 1972, Nuov.\ Cimm.\ Lett.\ {\bf 4,} 329.\\ Collins, J.\ C., 1974, Nucl.\ Phys.\ {\bf B 80,} 341.\\ Collins, J.\ C., 1984, {\sl Renormalization} (Cambridge University Press, Cambridge, UK).\\ Collins, J.\ C., A.\ Duncan, and S.\ D.\ Joglekar, 1977, Phys.\ Rev.\ {\bf D 16,} 438.\\ Collins, J.\ C., A.\ J.\ Macfarlane, 1974, Phys.\ Rev.\ {\bf D 10,} 1201.\\ Collins, J.\ C., and D.\ E.\ Soper, 1987, Annu.\ Rev.\ Nucl.\ Sci.\ {\bf 37,} 383.\\ Collins, J.\ C., D.\ E.\ Soper, and G.\ Sterman, 1983, in {\sl Proceedings of the 18th Rencontres de Moriond,} Edited by J.~Tran Thanh Van, p.\ 157.\\ Collins, J.\ C., D.\ E.\ Soper, and G.\ Sterman, 1984, Phys.\ Lett.\ {\bf B 134,} 263.\\ Collins, J.\ C., D.\ E.\ Soper, and G.\ Sterman, 1985, Nucl.\ Phys.\ {\bf B 261,} 104.\\ Collins, J.\ C., D.\ E.\ Soper, and G.\ Sterman, 1989, in {\sl Perturbative Quantum Chromodinamics,} Edited by A.\ H.\ Muller (World Scientific), p.\ 1.\\ Collins, J.\ C., F.\ Wilczek, and A.\ Zee, 1978, Phys.\ Rev.\ {\bf D 18,} 242.\\ Davydychev A.\ I., 1991, J.\ Math.\ Phys.\ {\bf 32,} 1052.\\ de~Rafael, E., and J.\ L.\ Rosner, 1974, Annals of Phys.\ {\bf 82,} 369.\\ de~R\'{u}jula, A., and H.\ Georgi, 1976, Phys.\ Rev.\ {\bf D 13,} 1296.\\ Delbourgo, R., and D.\ A.\ Akyeampong, 1974, Nuov.\ Cim.\ {\bf A 19,} 219.\\ de~Witt, B., 1967, Phys.\ Rev.\ {\bf 162,} 1195.\\ Diberder, F.\ L., and A.\ Pich, 1992a, Phys.\ Lett.\ {\bf B 286,} 147.\\ Diberder, F.\ L., and A.\ Pich, 1992b, Phys.\ Lett.\ {\bf B 289,} 165.\\ Dine, M., and J.\ Sapirstein, 1979, Phys.\ Rev.\ Lett.\ {\bf 43,} 668.\\ Dingle, R.\ B., 1973, {\it Asymptotic Expansions: Their derivation and Interpretation} (Academic Press, New York).\\ Drees, M., and K.\ Hikasa, 1990, Phys.\ Rev.\ {\bf D 41,} 1547.\\ Drell, S.\ D., and T.\ M.\ Yan, 1971, Ann.\ Phys.\ {\bf 66,} 578.\\ Duncan, A.\ H., {\it et al.}, 1993, Phys.\ Rev.\ Lett.\ {\bf 70,} 4159.\\ Efremov, A.\ V., and A.\ V.\ Radyushkin, 1980a, Teor.\ Mat.\ Fiz.\ {\bf 44,} 17 [Theor.\ Math.\ Phys.\ {\bf 44,} 573 (1980)].\\ Efremov, A. V., and A. V. Radyushkin, 1980b, Teor.\ Mat.\ Fiz.\ {\bf 44,} 157 [Theor.\ Math.\ Phys.\ {\bf 44,} 664 (1981)].\\ Ellis, J., M.\ Karliner, and M.\ Samuel, 1995, Phys.\ Rev.\ Lett.\\ Ellis, R.\ K., 1993, in {\sl Proceedings of the 7th 1992 Fermilab Meeting of the American Physical Society,} edited by C.\ H.\ Albright {\it et al.}, (World Scientific) p.\ 167.\\ Ellis, R.\ K., and W.\ J.\ Stirling, 1990, ``QCD AND COLLIDER PHYSICS,'' Fermilab preprint No.\ FERMILAB-Conf-90/164-T.\\ Faddeev, L.\ D, and U.\ N.\ Popov, 1967, Phys.\ Lett.\ {\bf B 25,} 29.\\ Faddeev, L.\ D, and A.\ A.\ Slavnov, 1980, {\it Gauge Fields: Introduction to Quantum Theory} (Benjamin, New York).\\ Feynman, R., 1963, Acta Phys.\ Polonica {\bf 26,} 697.\\ Feynman, R., 1969, Phys.\ Rev.\ Lett.\ {\bf 23,} 1415.\\ Feynman, R., 1972, {\it Photon Hadron Interactions} (Benjamin, New York).\\ Field, J.\ H., 1993, Ann.\ Phys.\ {\bf 226,} 209.\\ Fritzsch, H., M.\ Gell-Mann, and H.\ Leutwyler, 1973, Phys.\ Lett.\ {\bf B 47,} 365.\\ Furry, W., 1937, Phys.\ Rev.\ {\bf 51,} 125.\\ Gell-Mann, M., 1964, Phys.\ Lett., {\bf 8,} 214.\\ Gell-Mann, M., and F.\ Low, 1954, Phys.\ Rev.\ {\bf 95,} 1300.\\ Georgi, H., H.\ D.\ Politzer, 1976, Phys.\ Rev.\ {\bf D 14,} 1829.\\ Glashow, S.\ L., J.\ Iliopoulos, and L.\ Maiani, 1970, Phys.\ Rev.\ {\bf D 2,} 1285.\\ Gorishny, S.\ G., A.\ L.\ Kataev, S.\ A.\ Larin, 1986, Nuov.\ Cim.\ {\bf A 92,} 119.\\ Gorishny, S.\ G., A.\ L.\ Kataev, S.\ A.\ Larin, 1988, Phys.\ Lett.\ {\bf B 212,} 238.\\ Gorishny, S.\ G., A.\ L.\ Kataev, S.\ A.\ Larin, 1990, ``Four-loop QED $\beta$ function'' (private communications).\\ Gorishny, S.\ G., A.\ L.\ Kataev, S.\ A.\ Larin, 1991, Phys. Lett. {\bf B 259,} 144.\\ Gorishny, S.\ G., A.\ L.\ Kataev, S.\ A.\ Larin, and L.\ R.\ Surguladze, 1990, Mod.\ Phys.\ Lett.\ {\bf A 5,} 2703.\\ Gorishny, S.\ G., A.\ L.\ Kataev, S.\ A.\ Larin, and L.\ R.\ Surguladze, 1991a, Phys.\ Lett.\ {\bf B 256,} 81.\\ Gorishny, S.\ G., A.\ L.\ Kataev, S.\ A.\ Larin, and L.\ R.\ Surguladze, 1991b, Phys.\ Rev.\ {\bf D 43,} 1633.\\ Gorishny, S.\ G., A.\ L.\ Kataev, S.\ A.\ Larin, and L.\ R.\ Surguladze, 1991c, in {\sl Proceedings of the International Seminar ``QUARKS-90''} (Telavi, Georgia, USSR, May 1990) edited by V.~A.~Matveev {\it et al.}, (World Scientific) p.\ 194.\\ Gorishny, S.\ G., and S.\ A.\ Larin, 1987, Nucl.\ Phys.\ {\bf B 283,} 452.\\ Gorishny, S.\ G., S.\ A.\ Larin, L.\ R.\ Surguladze, and F.\ V.\ Tkachov, 1989, Comput.\ Phys.\ Commun.\ {\bf 55,} 381.\\ Gorishny, S.\ G., S.\ A.\ Larin, and F.\ V.\ Tkachov, 1983, Phys.\ Lett.\ {\bf B 124,} 217.\\ Greenberg, O.\ W., 1964, Phys.\ Rev.\ Lett.\ {\bf 13,} 598.\\ Gross, D., and F.\ Wilczek, 1973, Phys.\ Rev.\ Lett.\ {\bf 30,} 1343.\\ Grunberg, G., 1980, Phys.\ Lett.\ {\bf B 95,} 70.\\ Grunberg, G., 1982, Phys.\ Lett.\ {\bf B 110,} 501.\\ Grunberg, G., 1984, Phys.\ Rev.\ {\bf D 29,} 2315.\\ Grunberg, G., and A.\ L.\ Kataev, 1992, Phys.\ Lett.\ {\bf B 279,} 352.\\ Han, M.\ Y., and Y.\ Nambu, 1965, Phys.\ Rev.\ {\bf 139,} 1005.\\ Hearn, A.\ C., 1973, ``{\small REDUCE}, {\sl User's Manual} (University of Utah), Report No.\ UCP-19.\\ Hoang, A.\ H., M.\ Jezabek, J.\ H.\ K\"{u}hn, and T.\ Teubner, 1994, Phys.\ Lett.\ {\bf B 338,} 330.\\ Inami, T., and T.\ Kubota, 1981, Nucl.\ Phys.\ {\bf B 179,} 171.\\ Johnson, K., R.\ Willey, and M.\ Baker, 1967, Phys.\ Rev.\ {\bf 163,} 1699.\\ Johnson, K., and M.\ Baker, 1973, Phys.\ Rev.\ {\bf D 8,} 1110.\\ Kartvelishvili, V., and M.\ Margvelashvili, 1995, Phys.\ Lett.\ {\bf B 345,} 161.\\ Kataev, A.\ L., 1990, ``Next-next-to-leading perturbative QCD corrections: the current status of investigations,'' Montpellier Preprint No.\ PM/90-41.\\ Kataev, A.\ L., 1991, Nucl.\ Phys.\ B (Proc.\ Suppl.) {\bf A 23,} 72.\\ Kataev, A.\ L., and V.\ V.\ Starshenko, 1994, CERN Preprint No.\ TH.7400/94.\\ Kleinert, H., {\it et al.}, 1991, Phys.\ Lett.\ {\bf B 272,} 39.\\ Kniehl, B.\ A., 1990, Phys.\ Lett.\ {\bf B 237,} 127.\\ Kniehl, B.\ A., 1994a, Phys.\ Rep.\ {\bf 240,} 211.\\ Kniehl, B.\ A., 1994b, in {\sl Proceedings of the 1994 Tennessee International Symposium on Radiative Corrections: Status and Outlook,} (to be published); Bulletin Board: hep-ph/9410391.\\ Kniehl, B.\ A., 1995a, Phys.\ Lett.\ {\bf B 343,} 299.\\ Kniehl, B.\ A., 1995b, Int.\ J.\ Mod.\ Phys.\ {\bf A 10,} 443.\\ Kniehl, B.\ A., and J.\ H.\ K\"{u}hn, 1989, Phys.\ Lett.\ {\bf B 224,} 229.\\ Kniehl, B.\ A., and J.\ H.\ K\"{u}hn, 1990, Nucl.\ Phys.\ {\bf B 329,} 547.\\ Kotikov, A.\ V., 1991, Mod.\ Phys.\ Lett.\ {\bf A 6,} 677.\\ Kramer, G., and B.\ Lampe, 1988, Z.\ Phys.\ {\bf C 39,} 101.\\ Krasnikov, N.\ V., and A.\ A.\ Pivovarov, 1982, Phys.\ Lett.\ {\bf B 116,} 168.\\ Krasnikov, N.\ V., A.\ A.\ Pivovarov, and N.\ N.\ Tavkhelidze, 1983, Z.\ Phys.\ {\bf C 19,} 301.\\ Krasnikov, N.\ V., and N.\ N.\ Tavkhelidze, 1982, ``The contribution of instantons into cross-section of the $e^{+}e^{-}$ annihilation into hadrons'' Moscow Institute for Nuclear Research Preprint No.\ P-227.\\ Lam, C.\ S., and T.\ M.\ Yan, 1977, Phys.\ Rev.\ {\bf D 16,} 703.\\ Langacker, P., and L.\ Mingxing, and A.\ K.\ Mann, 1992, Rev.\ Mod.\ Phys.\ {\bf 64,} 87.\\ Lanin L.\ V., V.\ P.\ Spiridonov, and K.\ G.\ Chetyrkin, 1986, Yad.\ Fiz.\ {\bf 44,} 1374.\\ Larin S.\ A., 1993, Phys.\ Lett.\ {\bf B 303,} 113.\\ Lee, B.\ W., and J.\ Zinn-Justin, 1972, Phys.\ Rev.\ {\bf D 5,} 3121.\\ Lee, B.\ W., and J.\ Zinn-Justin, 1973, Phys.\ Rev.\ {\bf D 7,} 1049.\\ Le Guillou, J.\ C., and J.\ Zinn-Justin, 1990, Eds., {\sl Large-Order Behaviour of Perturbation Theory} (Elsevier Science Publishers B.\ V., North-Holland, Amsterdam).\\ Leibbrandt, G., 1975, Rev.\ Mod.\ Phys.\ {\bf 47,} 849.\\ Libby, S.\ B., and G.\ Sterman, 1978, Phys.\ Rev.\ {\bf D 18,} 3252.\\ Logunov, A.\ A., L.\ D.\ Soloviov, and A.\ N.\ Tavkhelidze, 1967, Phys.\ Lett.\ {\bf B 24,} 181.\\ Loladze, G.\ T., L.\ R.\ Surguladze, and F.\ V.\ Tkachov, 1984, Bull.\ Acad.\ Sci.\ Georgian SSR {\bf 116,} 509.\\ Loladze, G.\ T., L.\ R.\ Surguladze, and F.\ V.\ Tkachov, 1985, Phys.\ Lett.\ {\bf B 162,} 363.\\ Lovett-Turner, C.\ N., and C.\ J.\ Maxwell, 1994, Nucl.\ Phys.\ {\bf B 432,} 147.\\ Lu, H.\ J., and C.\ A.\ R.\ de Melo, 1991, Phys.\ Lett.\ {\bf B 273,} 260.\\ Mandelstam, S., 1968, Phys.\ Rev.\ {\bf 175,} 1580.\\ Marciano, W.\ J., 1975, Phys.\ Rev.\ {\bf D 12,} 3861.\\ Marciano, W.\ J., 1984, Phys.\ Rev.\ {\bf D 29,} 580.\\ Marciano, W.\ J., and H.\ Pagels, 1978, Phys.\ Rep.\ {\bf C 36,} 137.\\ Marciano, W.\ J., and A.\ Sirlin, 1988, Phys.\ Rev.\ Lett.\ {\bf 61,} 1815.\\ Marciano, W.\ J., 1991, {\sl Annu.\ Rev.\ Nucl.\ Sci.\ } {\bf 41,} 469.\\ Marciano, W.\ J., 1992, ``$\tau$ decays: a theoretical perspective,'' Brookhaven National Laboratory Preprint No.\ BNL-48179.\\ Marciano, W.\ J., 1993a, in {\sl Proceedings of the 7th 1992 Fermilab Meeting of the American Physical Society,} edited by C.\ H.\ Albright {\it et al.} (World Scientific), p.\ 185.\\ Marciano, W.\ J., 1993b, ``Standard Model Status'', Brookhaven National Laboratory Preprint No.\ BNL-48760.\\ Matveev, V.\ A., R.\ M.\ Muradyan, and A.\ N.\ Tavkhelidze, 1970, Fiz.\ Elem.\ Chastits At Yadra {\bf 1,} 91 [Sov.\ J.\ Part.\ Nucl.].\\ Matveev, V.\ A., R.\ M.\ Muradyan, and A.\ N.\ Tavkhelidze, 1972, Lett.\ Nuov.\ Cim.\ {\bf 5,} 907.\\ Matveev, V.\ A., R.\ M.\ Muradyan, and A.\ N.\ Tavkhelidze, 1973, Lett.\ Nuov.\ Cim.\ {\bf 7,} 719.\\ Mattingly, A.\ C., and P.\ M.\ Stevenson, 1992, Phys.\ Rev.\ Lett.\ {\bf 69,} 1320.\\ Mattingly, A.\ C., and P.\ M.\ Stevenson, 1994, Phys.\ Rev.\ {\bf D 49,} 437.\\ Maxwell, C.\ J., and J.\ A.\ Nicholls, 1990, Phys.\ Lett.\ {\bf B 236,} 63.\\ Miamoto, Y., 1965, Prog.\ Theor.\ Phys.\ Suppl.\ Extra {\bf 187}.\\ Monsay, E.\ and C.\ Rosenzweig, 1981, Phys.\ Rev.\ {\bf D 23,} 1217.\\ Mueller, A.\ H., 1978, Phys.\ Rev.\ {\bf D 18,} 3705.\\ Mueller, A.\ H., 1981, Phys.\ Rep.\ {\bf 73,} 237.\\ Mueller, A.\ H., 1992, in {\sl Proceedings of the Workshop QCD-Twenty Years Later,} edited by P.\ M.\ Zerwas, and H.\ A.\ Kastrup, (World Scientific) {\bf 1,} p.\ 162.\\ Muta, T., 1987, {\sl Foundations of Quantum Chromodinamics,} Lecture Notes in Physics Vol.\ 5 (World Scientific).\\ Nason, P., and M.\ Porrati, 1994, Nucl.\ Phys.\ {\bf B 421,} 518.\\ Narison, S., 1981a, Phys.\ Lett.\ {\bf B 104,} 485.\\ Narison, S., 1981b, Nucl.\ Phys.\ {\bf B 182,} 59.\\ Narison, S., 1982, Phys.\ Rep.\ {\bf 84,} 263.\\ Narison, S., 1986, ``QCD duality sum rules: introduction and some recent developments,'' CERN Preprint No.\ TH.4624/86.\\ Narison, S., 1994, ``$\alpha_s$ from tau decays'', CERN Preprint No.\ TH.7506/94.\\ Narison, S., and E.\ de Rafael, 1980, Nucl.\ Phys.\ {\bf B 169,} 253.\\ Narison, S., and E.\ de Rafael, 1981, Phys.\ Lett.\ {\bf B 103,} 57.\\ Narison, S., A.\ Pich, 1988, Phys.\ Lett.\ {\bf B 211,} 183.\\ Narison, S., and R.\ Tarrach, 1983, Phys.\ Lett.\ {\bf B 125,} 217.\\ Nielsen, N.\ K., 1977, Nucl.\ Phys.\ {\bf B 120,} 212.\\ Novikov, V.\ A., {\it et al.}, 1978, Phys.\ Rep.\ {\bf 41,} 1.\\ Novikov, V.\ A., M.\ A.\ Shifman, A.\ I.\ Vainshtein, and V.\ I.\ Zakharov, 1985, Nucl.\ Phys.\ {\bf B 249,} 445.\\ Ovsyannikov, L.\ V., 1956, Dok.\ Akad.\ Nauk SSSR {\bf 109,} 112 [Sov.\ Phys.\ Dokl.].\\ Pauli, W., and F.\ Villars, 1949, Rev.\ Mod.\ Phys.\ {\bf 21,} 433.\\ Pennington M.\ R., and G.\ G.\ Ross, 1982, Phys.\ Lett.\ {\bf B 102,} 167.\\ Peterman, A., 1979, Phys.\ Rep.\ {\bf 53,} 159.\\ Pich, A., 1990, ``Hadronic tau decays and QCD,'' CERN Preprint No.\ TH.5940/90.\\ Pich, A., 1991, in {\sl Heavy flavours,} edited by A.~J.~Buras and M.~Lindner (CERN, Geneva), p.\ 375.\\ Pich, A., 1992a, ``Tau physics and tau charm factories,'' CERN Preprint No.\ TH.6672/92.\\ Pich, A., 1992b, ``QCD predictions for the tau hadronic width and determination of $\alpha_s(M_{\tau}^2)$,'' CERN Preprint No.\ TH.6738/92.\\ Pich, A., 1994a, ``QCD predictions for the tau hadronic width: determination of $\alpha_s(M_{\tau}^2)$,'' Val\`{e}ncia University Preprint No.\ FTUV/94-71.\\ Pich, A., 1994b, ``The Standard Model of electroweak interactions,'' Val\`{e}ncia University Preprint No.\ FTUV/94-62.\\ Pivovarov, A.\ A., and L.\ R.\ Surguladze, 1991, Nucl.\ Phys.\ {\bf B 360,} 97.\\ Pivovarov, A.\ A., 1992a, Nuovo Cim.\ {\bf A 105,} 813.\\ Pivovarov, A.\ A., 1992b, Z.\ Phys.\ {\bf C 53,} 461.\\ Pivovarov, G.\ B., and F.\ V.\ Tkachov, 1988, Teor.\ Mat.\ Fiz.\ {\bf 77,} 51 [Theor Math. Phys. {\bf 77,} 1038 (1988)].\\ Pivovarov, G.\ B., and F.\ V.\ Tkachov, 1993, Int.\ J.\ Mod.\ Phys.\ {\bf A 8,} 2241.\\ Poggio, E., H.\ Quinn, and S.\ Weinberg, 1976, Phys.\ Rev.\ {\bf D 13,} 1958.\\ Politzer, H.\ D., 1973, Phys.\ Rev.\ Lett.\ {\bf 30,} 1346.\\ Pumplin, J., 1989, Phys.\ Rev.\ Lett.\ {\bf 63,} 576.\\ Pumplin, J., 1990, Phys.\ Rev.\ {\bf D 41,} 900.\\ Quigg C., 1983, {\sl Gauge Theories of the Strong, Weak and Electromagnetic interactions,} Frontiers In Physics 56 (Benjamin).\\ Raczka, P.\ A., 1995, Z.\ Phys.\ {\bf C 65,} 481.\\ Radyushkin, A.\ V., 1982, ``Optimized $\Lambda$ -parametrization for the QCD running coupling constant in spacelike and timelike regions,'' Dubna Joint Institute for Nuclear Research Preprint No.\ E2-82-159.\\ Radyushkin, A.\ V., 1983, Fiz. Elem. Chastits At Yadra {\bf 14,} 58 [Sov.J. Part. Nucl.].\\ Reinders, L.\ J., H.\ R.\ Rubinstein, and S.\ Yazaki, 1985, Phys.\ Rept.\ {\bf 127,} 1.\\ Reya, E., 1981, Phys.\ Rep.\ {\bf 69,} 195.\\ Rodrigo, G., and A.\ Santamaria, 1993, Phys.\ Lett.\ {\bf B 313,} 441.\\ Sakai, N., 1980, Phys.\ Rev.\ {\bf D 22,} 2220.\\ Salam, A., 1969, in {\sl Elementary particle Theory,} edited by N.\ Svartholm (Almqvist \& Wiksells, Stockholm), p.\ 367.\\ Samuel, M.\ A., and G.\ Li, 1994a, Int.\ J.\ Theor.\ Phys.\ {\bf 33,} 1461.\\ Samuel, M.\ A., and G.\ Li, 1994b, Phys.\ Lett.\ {\bf B 331,} 114.\\ Samuel, M.\ A., and G.\ Li, 1994c, Int.\ J.\ Theor.\ Phys.\ {\bf 33,} 2207.\\ Samuel, M.\ A., G.\ Li, and E.\ Steinfelds, 1994a, ``On estimating perturbative coefficients in quantum field theory and statistical physics,'' Oklahoma State University Preprint No.\ RN-278.\\ Samuel, M.\ A., G.\ Li, and E.\ Steinfelds, 1994b, Phys.\ Lett.\ {\bf B 323,} 188.\\ Samuel, M.\ A., G.\ Li, and E.\ Steinfelds, 1994c, Phys.\ Rev.\ {\bf D 48,} 869.\\ Samuel, M.\ A., and L.\ R.\ Surguladze, 1991, Phys.\ Rev.\ {\bf D 44,} 1602.\\ Schilcher, K., and M.\ D.\ Tran, 1984, Phys.\ Rev.\ {\bf D 29,} 570.\\ Shankar, R., 1977, Phys.\ Rev.\ {\bf D 15,} 755.\\ Shifman, M.\ A., 1992, {\sl Vacuum Structure and QCD Sum Rules,} (Elsevier Science Publishers).\\ Shifman, M.\ A., A.\ I.\ Vainshtein, and V.\ I.\ Zakharov, 1979, Nucl.\ Phys.\ {\bf B 147,} 385.\\ Shirkov, D.\ V., 1980, ``Three loop approximation for running coupling constant in Quantum Chromodynamics,'' Dubna Joint Institute for Nuclear Research Preprint No.\ E2-80-609.\\ Shirkov, D.\ V., 1992, ``Historical remarks on the renormalization group,'' Max Planck Institute Preprint No.\ MPI-PAE/PTh 55/92.\\ Sirlin, A., 1993a, ``Universality of the weak interactions,'' New York University Preprint No.\ 93-0526.\\ Sirlin, A., 1993b, ``Status of the standard electroweak model,'' New York University Preprint No.\ NYU-TH-93-06-04.\\ Smirnov V.\ A., 1990, Commun.\ Math.\ Phys.\ {\bf 134,} 109.\\ Smirnov V.\ A., 1991, {\sl Renormalization and Asymptotic Expansions} (Birkhauser).\\ Smirnov, V.\ A., and K.\ G.\ Chetyrkin, 1985, Teor. Mat. Fiz. {\bf 63,} 208 [Theor.\ Math.\ Phys.\ {\bf 63,} 462 (1985)].\\ Soper, D.\ E., 1995, in {\sl Proceedings of the XXXth Rencontres de Moriond ``QCD and High Energy Interactions''} (Les Arcs, France).\\ Soper, D.\ E., and L.\ R.\ Surguladze, 1994, Phys.\ Rev.\ Lett.\ {\bf 73,} 2958.\\ Soper, D.\ E., and L.\ R.\ Surguladze, 1995, in {\sl Proceedings of the XXXth Rencontres de Moriond ``QCD and High Energy Interactions''} (Les Arcs, France).\\ Soper, D.\ E., and L.\ R.\ Surguladze, 1995 (in preparation).\\ Speer, E.\ R., 1974, J.\ Math.\ Phys.\ {\bf 15,} 1.\\ Spiridonov, V.\ P., 1984, ``Anomalous dimension of $G^2$ and $\beta$ function,'' Moscow Institute for Nuclear Research Preprint No.\ P-378.\\ Spiridonov, V.\ P., 1987, Yad. Fiz. {\bf 46,} 302 [Sov.\ J.\ Nucl.\ Phys.].\\ Stevenson, P.\ M., 1981a, Phys.\ Lett.\ {\bf B 100,} 61.\\ Stevenson, P.\ M., 1981b, Phys.\ Rev.\ {\bf D 23,} 2916.\\ Stevenson, P.\ M., 1982, Nucl.\ Phys.\ {\bf B 203,} 472.\\ Stevenson, P.\ M., 1984, Nucl.\ Phys.\ {\bf B 231,} 65.\\ Stevenson, P.\ M., 1992, ``Response to Brodsky and Lu's Letter: On the selfconsistency of scale setting methods,'' Rice University Preprint No.\ DOE-ER-40717-2; Bulletin Board: hep-ph/9211327.\\ Stevenson, P.\ M., 1994, (private communication).\\ Strubbe, H., 1974, Comput.\ Phys.\ Commun.\ {\bf 8,} 1.\\ Stueckelberg, E.\ C.\ G., and A.\ Peterman, 1953, Helv.\ Phys.\ Acta {\bf 26,} 499.\\ Surguladze, L.\ R., 1989a, ``$O(m^2)$ contributions to correlators of quark currents: three-loop approximation,'' Moscow Institute for Nuclear Research Preprint No.\ P-639.\\ Surguladze, L.\ R., 1989b, ``Structure of the program for multiloop calculations in quantum field theory on the {\small SCHOONSCHIP} system,'' Moscow Institute for Nuclear Research Preprint No.\ P-643.\\ Surguladze, L.\ R., 1989c, ``Program {\small MINCER} in Four-loop calculations'' (unpublished).\\ Surguladze, L.\ R., 1989d, Yad.\ Fiz.\ {\bf 50,} 604 [Sov.\ J.\ Nucl.\ Phys.\ {bf 50,} 372 (1989)].\\ Surguladze, L.\ R., 1990, ``Four-loop QED $\beta$ function'' (unpublished).\\ Surguladze, L.\ R., 1992, ``A program for analytical perturbative calculations in high energy physics up to four loops for the {\small FORM} system,'' Fermilab Preprint No.\ FERMILAB-PUB 92/191-T.\\ Surguladze, L.\ R., 1994a, Phys.\ Lett.\ {\bf B 338,} 229.\\ Surguladze, L.\ R., 1994b, Phys.\ Lett.\ {\bf B 341,} 60.\\ Surguladze, L.\ R., 1994c, ``Quark mass corrections to the Z boson decay rates,'' University of Oregon Preprint No.\ OITS-554.\\ Surguladze, L.\ R., 1994d, Int.\ J.\ Mod.\ Phys.\ {\bf C 5,} 1089.\\ Surguladze, L.\ R., and F.\ V.\ Tkachov, 1986, ``Three-loop coefficient functions of gluon and quark condensates in QCD sum rules for light mesons,'' Moscow Institute for Nuclear Research Preprint No.\ P-501.\\ Surguladze, L.\ R., and F.\ V.\ Tkachov, 1988, Teor.\ Mat.\ Fiz.\ {\bf 75,} 245 [Theor.\ Math.\ Phys.\ {\bf 75,} 502 (1988)].\\ Surguladze, L.\ R., and F.\ V.\ Tkachov, 1989a, Comp.\ Phys.\ Commun.\ {\bf 55,} 205.\\ Surguladze, L.\ R., and F.\ V.\ Tkachov, 1989b, Mod.\ Phys.\ Lett.\ {\bf A 4,} 765.\\ Surguladze, L.\ R., and F.\ V.\ Tkachov, 1990, Nucl.\ Phys.\ {\bf B 331,} 35.\\ Surguladze, L.\ R., and M.\ A.\ Samuel, 1991a, in {\sl Proceedings of the International Conference Beyond the Standard Model II} (Norman, OK, USA, 1990), edited by K.\ Milton, R.\ Kantowski, and M.\ A.\ Samuel (World Scientific), p.\ 206.\\ Surguladze, L. R., and M. A. Samuel, 1991b, Phys. Rev. Lett. {\bf 66,} 560.\\ Surguladze, L.\ R., and M.\ A.\ Samuel, 1992a, ``On West's asymptotic estimate of perturbative coefficients of R(s) in $e^{+}e^{-}$ annihilation,'' Oklahoma State University Preprint No.\ RN-268A.\\ Surguladze, L.\ R., and M.\ A.\ Samuel, 1992b, ``Four-loop perturbative calculations of $\sigma_{\mbox{\scriptsize tot}}(e^{+}e^{-}\rightarrow \mbox{hadrons})$, $\Gamma(\tau\rightarrow \nu_{\tau}+\mbox{hadrons})$ and QED $\beta$ function,'' Fermilab Preprint No.\ FERMILAB-PUB 92/192-T.\\ Surguladze, L.\ R., and M.\ A.\ Samuel, 1993, Phys.\ Lett.\ {\bf B 309,} 157.\\ Symanzik, K., 1970, Commun.\ Math.\ Phys.\ {\bf 18,} 227.\\ Symanzik, K., 1971, Commun.\ Math.\ Phys.\ {\bf 23,} 49.\\ Tarasov, O.\ V., 1982, ``Anomalous dimensions of quark masses in three-loop approximation,'' Dubna Joint Institute for Nuclear Research Preprint No.\ JINR-P2-82-900.\\ Tarasov, O.\ V., A.\ A.\ Vladimirov, and A.\ Yu.\ Zharkov, 1980, Phys.\ Lett.\ {\bf B 93,} 429.\\ Tarrach, R., 1982, Nucl.\ Phys.\ {\bf B 196,} 45.\\ Tavkhelidze, A.\ N., 1965, Lect.\ High Energy Phys.\ Elem.\ Particles (Vienna).\\ Tavkhelidze, A.\ N., 1994, ``Color, colored quarks, Quantum Chromodynamics,'' Dubna Joint Institute for Nuclear Research Preprint No.\ JINR-E2-94-372.\\ t 'Hooft, G., 1971, Nucl.\ Phys.\ {\bf B 33,} 173.\\ t 'Hooft, G., 1973, Nucl.\ Phys.\ {\bf B 61,} 455.\\ t 'Hooft, G., and M.\ Veltman, 1972, Nucl.\ Phys.\ {\bf B 44,} 189.\\ t 'Hooft, G., and M.\ Veltman, 1973, ``Diagrammar,'' CERN report.\\ Tkachov, F.\ V., 1981, Phys.\ Lett.\ {\bf B 100,} 65.\\ Tkachov, F.\ V., 1983a, Teor.\ Mat.\ Fiz.\ {\bf 56,} 350 [Theor.\ Math.\ Phys.\ {\bf 56,} 866 (1983)].\\ Tkachov, F.\ V., 1983b, Phys.\ Lett.\ {\bf B 124,} 212.\\ Tkachov, F.\ V., 1983c, Phys.\ Lett.\ {\bf B 125,} 85.\\ Tkachov, F.\ V., 1991, Fermilab Preprint No.\ FERMILAB-PUB-91/347-T.\\ Tkachov, F.\ V., 1993, Int.\ Journ.\ Mod.\ Phys.\ {\bf A 8,} 2047.\\ Trueman, T.\ L., 1979, Phys.\ Lett.\ {\bf B 88,} 331.\\ Tsai, Y.\ S., 1971, Phys.\ Rev.\ {\bf D 4,} 2821.\\ Vainshtein, A.\ I., and V.\ I.\ Zakharov, 1994, Phys.\ Rev.\ Lett.\ {\bf 73,} 1207.\\ Veltman, M., 1967, {\sl {\small SCHOONSCHIP}, A CDC 6600 program for symbolic evaluation of algebraic expressions} (CERN).\\ Veltman, M., 1991, {\sl {\small SCHOONSCHIP}, A program for symbol handling} (Michigan).\\ Vermaseren, J.\ A.\ M., 1989, {\sl {\small FORM}, User's Manual} (NIKHEP, Amsterdam).\\ Vladimirov, A.\ A., 1978, Teor.\ Mat.\ Fiz.\ {\bf 36,} 271 [Theor.\ Math.\ Phys.\ {\bf 36,} 732 (1979)].\\ Vladimirov, A.\ A., 1980, Teor.\ Mat.\ Fiz.\ {\bf 43,} 280 [Theor.\ Math.\ Phys.\ {\bf 43,} 417 (1980)].\\ Ward, J.\ C., 1950, Phys.\ Rev.\ {\bf 78,} 182.\\ Weinberg, S., 1967, Phys.\ Rev.\ Lett.\ {\bf 19,} 1264.\\ Weinberg, S., 1973, Phys.\ Rev.\ {\bf D 8,} 3497.\\ West, G.\ B., 1991, Phys.\ Rev.\ Lett.\ {\bf 67,} 1388.\\ Wetzel, W., and W.\ Bernreuther, 1981, Phys.\ Rev.\ {\bf D 24,} 2724.\\ Wilson, K.\ G., 1969, Phys.\ Rev.\ {\bf 179,} 1499.\\ Yang, C.\ N., and R.\ L.\ Mills, 1954, Phys.\ Rev.\ {\bf 96,} 191.\\ Yang, C.\ N., 1969, in {\sl High Energy Collisions} (Gordon\& Breach, NY), p.\ 509.\\ Yennie, D.\ R., S.\ C.\ Frautschi, and H.\ Suura, 1961, Ann.\ Phys.\ {\bf 13,} 379.\\ Yndurain, F.\ J., 1983, {\sl QCD: an Introduction to the Theory of Quarks an Gluons} (Springer Verlag).\\ Zakharov, V.\ I., 1992, Nucl.\ Phys.\ {\bf B 385,} 452.\\ Zweig, G., 1964, ``An SU(3) model for strong interaction symmetry and its breaking,'' CERN Preprint No.\ TH.412. \end{document}
\section{Introduction}\label{sec:intro} It has been known for some time that a majority of main-sequence stars are in binary systems (e.g.\ Abt 1983, Duquennoy \& Mayor 1991). Recently, systematic surveys of pre--main-sequence\ (PMS) stars have shown that their binary frequency is at least as high as, and perhaps significantly higher than, that of main-sequence stars (Mathieu, Walter, \& Myers 1989, Simon et al.\ 1992, 1995, Ghez, Neugebauer, \& Matthews 1993, Leinert et al.\ 1993, Reipurth \& Zinnecker 1993). Thus, the most likely outcome of the star formation process is a binary system and understanding binary formation is vital to understanding star formation. Infrared and millimeter observations have shown that many young stars have excess long-wavelength emission, indicative of circumstellar material. Although this material is unresolved in most observations, there is mounting evidence that in most cases it lies in circumstellar or circumbinary disks (see e.g.\ Beckwith \& Sargent 1993, Basri \& Bertout 1993, and Sargent 1995 for reviews). Indeed, much of a star's mass may be built up by accretion of material through these disks. Thus, disks provide a fossil record of the star formation process, and study of disk properties may reveal clues about star formation. Observations suggest that the structure of a disk is affected by the presence of an embedded binary. Beckwith et al.\ (1990, hereafter BSCG) measured 1300 \micron\ continuum emission from 86 PMS stars in Taurus-Auriga (Tau-Aur) to search for disks. They found that very few of the binaries in their sample with projected separations less than 100 AU had detectable 1300 \micron\ emission while strong emission was common in wider binaries (see also Beckwith \& Sargent 1993). Jensen, Mathieu, \& Fuller (1994, hereafter Paper I) used a larger sample of binaries in Tau-Aur and more sensitive 800 \micron\ observations of the systems with the smallest projected separations to further explore this result. We found that the fluxes from binaries with projected separations between 1 and 50--100 AU were significantly lower than those from wider binaries or single stars. On the other hand, the flux distributions of the two latter groups are statistically indistinguishable. Osterloh \& Beckwith (1995) made additional 1300 \micron\ observations of PMS stars in Tau-Aur and did detect some close binaries. Nonetheless, they found the difference in flux distributions between close binaries and wide binaries or single stars to remain at a statistically significant level. There are a variety of theoretical predictions about the influence of binaries on their associated disks. A binary companion embedded in a disk is expected to rapidly clear a gap, isolating distinct circumstellar and circumbinary disks. Such a gap may inhibit the transfer of material from circumbinary to circumstellar disks (e.g.\ Artymowicz et al.\ 1991, but see Artymowicz \& Lubow 1994). If so, continued accretion from the circumstellar disks onto the stars may exhaust the reservoir of circumstellar disk material more quickly than in single stars (Clarke 1992). Indeed binary companions may accelerate the depletion. Ostriker, Shu, \& Adams (1992) found that for separations less than 100 AU a companion can excite density waves in a circumstellar disk, causing an enhanced accretion rate onto the central star. At the same time, the inhibition of accretion could increase the lifetimes of circumbinary disks (Clarke 1992). Pringle (1991) found that the transfer of angular momentum from the binary orbit to the circumbinary disk pushes disk material to larger radii, resulting in a reduction in surface density and increase in size of the circumbinary disk. We note that most of these predictions remain largely untested by current observations. We have undertaken a submillimeter survey of young binaries in the Scorpius-Ophiuchus (Sco-Oph) star-forming region. In this paper, we present 800 \micron\ observations of most known Sco-Oph PMS binaries with projected separations $a_p$ less than 150 AU\null. We combine these data with other Sco-Oph submillimeter data from Andr\'e \& Montmerle (1994) and find a dependence of submillimeter flux on binary separation in a sample that is independent of that used by BSCG, Paper I, or Osterloh \& Beckwith (1995). We compare the submillimeter data from Sco-Oph and Tau-Aur with submillimeter fluxes predicted from models of disks with gaps and show that gap clearing alone may be sufficient to explain the low submillimeter fluxes from the close binaries. Most PMS binaries at all separations were detected at 60 \micron\ by IRAS, indicating the presence of circumstellar disks. We derive lower limits on circumstellar disk masses from the 60 \micron\ emission. Finally, we use submillimeter fluxes to place upper limits of 0.005 \hbox{${M}_{\sun}$}\ on circumbinary disk masses for typical binaries with $a_p$ between a few AU and 100 AU. \section{Observations} The target binaries (see \S\ref{section:sample}) were observed using the James Clerk Maxwell Telescope (JCMT)\footnote{The James Clerk Maxwell Telescope is operated by the Royal Observatories on behalf of the Particle Physics and Astronomy Research Council of the United Kingdom, the Netherlands Organization for Scientific Research, and the National Research Council of Canada.} on 1992 February 24--27, 1992 March 2--4, and 1993 February 26 -- March 1. The JCMT facility $^3$He cooled bolometer, UKT14, was used for the observations. All observations were made using a 65-mm aperture which resulted in a $\sim19\arcsec$ beam width (FWHM) at all wavelengths. All the observations were made using standard chopping and beam switching with synchronous detection of the signal. The secondary chopper was switched 60\arcsec\ in azimuth at a rate of 7.8 Hz. During the 1992 observing sessions the atmosphere was usually stable and transparent. Measurements from a tipping radiometer indicated zenith optical depths of typically 0.03 to 0.04 at 225 GHz. The weather during the 1993 run was very stable and the atmosphere transparent. The zenith optical depth at 225 GHz was approximately constant at $\sim0.08$ during the first night and between 0.04 and 0.05 for the remaining nights. During both runs standard sources of known flux were observed every 15 to 30 minutes in order to monitor the sky opacity. The pointing of the telescope was also checked and corrected regularly during the observations by making five-point measurements on bright sources. Typically the actual pointing was found to be within $\sim3$\arcsec\ of the nominal pointing. Very occasional shifts as large as 6\arcsec\ were seen. Our survey observations were made at 800 \micron. Because thermal dust emission decreases with wavelength from 800 to 1300 \micron\ more quickly than the atmospheric transmission improves (in good weather), at the JCMT 800 \micron\ observations are more sensitive for detecting weak dust emission. In addition, non-thermal emission (e.g.\ free-free or gyrosynchrotron) increases in strength with increasing wavelength, and thus 800 \micron\ observations are less likely to be contaminated by this non-disk emission. The program sources were initially observed for $\sim 1000$ s at 800 \micron. This resulted in a typical 1$\sigma$ noise level of 15--20 mJy. Most sources were observed twice, particularly those for which the noise level of the first observation was higher than the typical level of sources which were marginally detected. Observations at 1100 \micron\ and 450 \micron, and occasionally 350 \micron, were made of sources which were strong detections at 800 \micron. Observations of the standard sources at different airmasses were used to determine the telescope sensitivity at 1100 \micron, 800 \micron, 450 \micron, and 350 \micron. Given the telescope sensitivity, the standard source observations provided an estimate of the sky opacity as a function of time throughout each night. For both observing sessions, the variation of the sky opacity with time derived from these observations very closely tracked the opacity derived from the 225 GHz radiometer. The program sources were calibrated using the mean sky opacity derived from standard sources observed before and after the target source. The scatter in the zenith opacity about a linear fit versus time for each night has been used as an estimate of the uncertainty in the flux calibration. Each observation consisted of a series of 10-second samples. To remove noise spikes from the data, the mean of the samples in each integration was calculated and those samples which deviated from the mean by more than three standard deviations were removed and the mean was recalculated. The measured fluxes or $3\sigma$ upper limits are given in Table \ref{table:jcmt_data}. The quoted uncertainties are a combination of the flux calibration uncertainty discussed above and the photon noise derived from the standard deviation of the individual 10-second samples within each observation. \section{Sample selection}\label{section:sample} We chose binaries for our sample from among members of the Upper Scorpius subgroup of the Sco OB2 association and from the $\rho$ Oph cloud complex. These stars lie within the approximate boundaries of right ascension 15$^{\rm h}$10$^{\rm m}$--16$^{\rm h}$40$^{\rm m}$ and declination $-5\deg$--$-35\deg$ (Blauuw 1978). We will refer to this area as ``Sco-Oph''. We adopt distances of 125 pc for the $\rho$ Oph clouds and 160 pc for Upper Sco (de Geus, de Zeeuw, \& Lub 1989). Surveys which have searched for binaries in Sco-Oph include those of Mathieu et al.\ (1989), Ghez et al.\ (1993), Reipurth \& Zinnecker (1993), and Simon et al.\ (1992, 1995). These surveys vary greatly in their target lists and the range of binary separations to which they are sensitive. The characteristics of the surveys are summarized in Table \ref{table:binary_surveys}. In addition, several PMS binaries have been found by observations that were not part of surveys. AK Sco was discovered to be a spectroscopic binary by Andersen et al.\ (1989). WL 2 and WL 20 were both found to be binaries by infrared imaging (Rieke, Ashok, \& Boyle 1989, Barsony et al.\ 1989). VV Sco was noted as a visual binary by Gregorio-Hetem et al.\ (1992). DoAr 51 was found to be a binary by Koresko (1995) using speckle interferometry. Currently only spectroscopic observations can detect binaries with separations less than $\sim$ 1 AU at the distance of Sco-Oph. All but one (AK Sco) of the known spectroscopic binaries in Sco-Oph were found in a radial-velocity survey of x-ray selected stars in Sco-Oph (Mathieu et al.\ 1989). Independent of multiplicity, most of the x-ray selected stars surveyed by Mathieu et al.\ (1989) do not have infrared excesses or strong H$\alpha$ emission (Walter et al.\ 1994), the classical signatures of disks. Thus the binaries discovered in that survey may not be representative of PMS spectroscopic binaries in general, particularly in terms of the disk properties with which we are concerned here. In contrast to the spectroscopic observations of Mathieu et al.\ (1989), the other surveys for binaries in Sco-Oph have drawn their target lists primarily from lists of H$\alpha$- and infrared-selected PMS stars. Ghez et al.\ (1993) and Simon et al.\ (1995) also include some x-ray selected PMS stars in their target lists. The targets of these surveys were not initially discovered to be young stars using a uniform set of selection criteria, and thus the combined sample of Sco-Oph binaries from these surveys is not a well-defined, uniformly-selected PMS binary sample. However, the diversity of targets of these surveys reflects the variety of techniques that have been used to search for young stars. Thus, while the sample of binaries discovered by these surveys may not be uniformly defined, it should be reasonably representative of the known population of PMS stars in Sco-Oph in general. Because the spectroscopic binaries are drawn from a distinctly different sample of PMS stars than are the other binaries, in this paper we limit our analysis to binaries with projected separations greater than 1 AU\null. This sample should be neither more nor less biased toward the presence of disks than the known Sco-Oph PMS population as a whole. For these binaries, we present a quantitative analysis of the distribution of submillimeter fluxes with projected separation and discuss the implications of our data for binary-disk interactions. We will discuss the submillimeter fluxes and spectral energy distributions of binaries with projected separations less than 1 AU in a later paper. Simon et al.\ (1992, 1995) argue that for separations less than 10\arcsec, contamination of the binary sample from chance superposition of stars should be negligible. Thus, we adopt this as the separation upper bound for our sample. We have taken projected separations or projected semimajor axes for Sco-Oph binaries from the binary surveys discussed above. The other primary requirement for inclusion in our sample is availability of an 800 \micron\ or 1300 \micron\ flux measurement. (For convenience, we will refer to either of these as ``submillimeter fluxes.'') Submillimeter fluxes or flux limits are from this work and Andr\'e \& Montmerle (1994, hereafter AM), which together cover about 90\% of the currently known PMS binaries in Sco-Oph. Our goal is to explore the influence of multiplicity on the properties of circumstellar and circumbinary disks. In order to avoid confusion introduced by other factors that may also influence submillimeter emission, we have further constrained our sample. Our sample is limited to PMS stars with spectral types of F and later so as to reduce the range of stellar mass, luminosity, and temperature. We have also excluded Class I sources as defined in the infrared classification scheme of Lada (1987). Terebey, Chandler, \& Andr\'e (1993) suggest that a significant fraction of the 1300 \micron\ emission from Class I sources arises in an envelope rather than a disk; this is consistent with the fact that Class I sources have significantly higher 1300 \micron\ fluxes on average than Class II or Class III sources and that Class I sources tend to be extended whereas Class II sources are unresolved (AM). Class I sources are also heavily embedded and presumably younger than Class II or III sources (AM). In fact, the spectral type and infrared class criteria remove only a few systems from our sample. After applying these selection criteria, we obtain a sample of 30\ multiple systems in Sco-Oph. This is comparable to the 42 systems in Tau-Aur studied in Paper I\null. With the addition of data from Osterloh \& Beckwith (1995), the present Tau-Aur sample includes 55\ multiple systems. We also analyze this expanded sample below. In addition to the binaries, we adopt a sample of single stars for comparison. Single stars provide a ``control'' sample, revealing submillimeter fluxes produced in the absence of any stellar-mass companion. We take our single-star sample from the targets observed by AM which are not known to have stellar companions within 10\arcsec. We apply the same spectral type and IR class criteria described above. From the $\rho$ Oph region, we adopt only the sources that AM have marked with a ``\#'' in their Table 1, since these are all confirmed PMS stars which have been observed at a uniform sensitivity. In addition, we adopt the other single PMS stars which AM observed that lie outside $\rho$ Oph but within Sco-Oph. This yields a total of 47\ stars. We note that this definition of our single star sample is less restrictive than in Paper I for the Tau-Aur sample, where we adopted as single only those stars which had been surveyed using speckle interferometry or lunar occultation and found not to have companions. However, the surveys for binaries in Sco-Oph have been less comprehensive and less uniform than in Tau-Aur; applying such a criterion to Sco-Oph would yield a prohibitively small sample of 15 stars. Thus, here we adopt as single those stars not known to have companions, whether or not they have been surveyed with high-resolution techniques. To estimate the contamination from unresolved companions in our single-star sample, we note that of 47\ stars in the ``single'' sample, 15 have been surveyed using high-resolution techniques, leaving 32 of unknown multiplicity. Adopting a multiplicity frequency of 50\% (Simon et al.\ 1995 find 48\% for Tau-Aur and Oph), we then expect $\sim 16$ binaries out of 47\ stars in the sample. Wide companions ($a_p \gtrsim 1\arcsec$) would be more likely to have been previously discovered, so the contamination is likely to be predominantly close binaries. Thus the expected effect of contamination of the single sample by undetected binaries would be to weaken the statistical significance of any differences between the flux distributions of the close binaries and single stars. \section{The dependence of submillimeter flux on binary separation} \subsection{Combining 1300 \micron\ and 800 \micron\ flux measurements} The bulk of the submillimeter and millimeter continuum flux measurements of PMS stars in the literature are contained in the surveys of BSCG, Osterloh \& Beckwith (1995) (both including Tau-Aur), and AM (Sco-Oph). All of these surveys were made at a wavelength of 1300 \micron\ while our observations were made at 800 \micron. Combining these measurements raises the question of how to compare fluxes measured at 1300 \micron\ with those measured at 800 \micron. Because of the existence of a large number of 1300 \micron\ flux measurements of PMS stars in the literature, we have chosen to scale our 800 \micron\ fluxes or flux limits to 1300 \micron\ for comparison with existing data. Submillimeter and millimeter fluxes from dust around PMS stars are dominated by emission on the Rayleigh-Jeans tail of the Planck function, and thus if the emission is optically thick the flux density $F_\nu$ should vary with frequency as $\nu^2$. If the emission is optically thin and the submillimeter dust opacity varies as $\kappa_{\nu} \propto \nu^\beta$, then the flux should vary as $\nu^{2+\beta}$. Since dust opacity is thought to increase with increasing frequency (i.e.\ $\beta$ is positive; Beckwith \& Sargent 1991, Pollack et al.\ 1994, Agladze et al.\ 1994), then the variation of submillimeter and millimeter thermal continuum emission with frequency should be {\em at least\/} as steep as $F_\nu \propto \nu^2$. In fact, observations of PMS stars at multiple millimeter wavelengths bear this out. Beckwith \& Sargent (1991) observed 25 PMS stars and found that 19 out of 25 have millimeter spectral slopes greater than or equal to $\nu^2$ within their 1$\sigma$ uncertainties. Thus, for systems with upper limits on their 800 \micron\ fluxes we calculate 1300 \micron\ upper limits by multiplying the 800 \micron\ flux limits by $(\nu_{1300}/ \nu_{800})^2 = 0.38$. Since the true spectral slope may be steeper than $\nu^2$, this scaling provides a conservative but secure upper limit on the 1300 \micron\ flux. For systems with upper limits at both 800 and 1300 \micron, the measured 1300 \micron\ limit is compared to the 800 \micron\ limit scaled to 1300 \micron, and the lower of the two values is used. One system in our Sco-Oph sample was detected at 800 \micron\ but not 1300 \micron. Haro 1-4 was not detected by AM with a 1300-\micron\ flux limit of $F_\nu < 50 $ mJy. This limit is consistent with our 800 \micron\ detection of $F_\nu = 58 $ mJy. To estimate the 1300 \micron\ flux of Haro 1-4, we extrapolated the 800-\micron\ flux to 1300 \micron\ assuming a spectral slope of $\nu^2$, the most common slope found by Beckwith \& Sargent (1991). Nonetheless, the unknown spectral slope leads to an uncertain 1300 \micron\ flux. However, tests using different slopes to determine the flux of Haro 1-4, or leaving it out of the sample altogether, show that the effect of this uncertainty on our results is negligible. A few binary systems do not show the variation of submillimeter flux with wavelength expected for thermal emission from dust. We treat these on a case-by-case basis. V773 Tau was detected at 1300 \micron\ by BSCG with a flux of $42 \pm 6$ mJy and by Osterloh \& Beckwith (1995) with a flux of $24 \pm 4$ mJy. However, we did not detect it at 800 \micron\ with a $3\sigma$ upper limit of 29 mJy (Paper I). The apparent variability at 1300 \micron, the 800--1300 \micron\ spectral slope much flatter than $\nu^2$, and the strong non-thermal centimeter wavelength emission (Phillips, Lonsdale, \& Feigelson 1991) suggest that at least some of the 1300 \micron\ flux from V773 Tau is produced by a non-thermal process and is not emission from circumstellar dust. As in Paper I, we proceed on the assumption that the 1300 \micron\ emission is non-thermal, and we use the 800 \micron\ limit from Paper I to derive an upper limit on the 1300 \micron\ disk emission. Two binaries in Sco-Oph (SR 9 and DoAr 24E) show a similar disagreement of 800 \micron\ and 1300 \micron\ fluxes. We did not detect SR 9 at 800 \micron\ with a $3\sigma$ upper limit of 28 mJy. Scaling this to 1300 \micron\ by $\nu^2$ yields an upper limit of 11 mJy, whereas AM list it as a weak detection ($\sim 4\sigma$) at 1300 \micron\ with a flux of 15 mJy. However, in this case there is no independent evidence of non-thermal emission; SR 9 was undetected at 6 cm (Andr\'e et al.\ 1990). SR 9 may be similar to the few systems found by Beckwith \& Sargent (1991) to have submillimeter spectral slopes slightly flatter than $\nu^2$ (i.e.\ $-1 < \beta < 0$). Because of the lack of other evidence for non-thermal contributions to the millimeter flux and the relatively modest deviation from expected dust-emission spectral slope, for SR 9 we adopt the 1300 \micron\ flux measured by AM for our analysis. DoAr 24E, however, shows a larger disagreement between 800 and 1300 \micron\ fluxes. We measured a flux of $37\pm 8$ mJy at 800 \micron, while AM measured 65 mJy at 1300 \micron. These measurements imply $\beta \approx -3.2$, well outside the range of $\beta$ values found by Beckwith \& Sargent (1991). Reipurth \& Zinnecker (1993) report a 1300 \micron\ flux from DoAr 24E of $\sim 20$ mJy. Although no radio continuum measurements of DoAr 24E are known to us, its unusual submillimeter spectral slope and apparent variability (as in the case of V773 Tau) suggest a non-thermal contribution to its 1300 \micron\ flux. Thus, we adopt our 800 \micron\ flux scaled to 1300 \micron\ by $\nu^2$ as the best estimate of the disk contribution to submillimeter emission for the analysis below. Finally, we emphasize that, given the size of our sample, the particular choice of flux for these three systems (or even their inclusion or omission) does not affect the conclusions of our analysis. \subsection{Assignment of flux in unresolved triple and quadruple systems} Five systems in our Sco-Oph sample each comprise three or four stars that lie within the JCMT beam: V853 Oph, SR 24, 155913$-$2233, 162814$-$2427 (ROX 42C), and 162819$-$2423 (ROX 43A/B). In the Tau-Aur binary sample, there are six systems which are triples with all components within 10\arcsec\ of each other: HP Tau/G3, V807 Tau, HV Tau, RW Aur, UZ Tau, and UX Tau. In the absence of spatially-resolved flux measurements of the individual components of such higher-order multiple systems, it is unknown whether any detected flux should be associated with the binary, the tertiary, or both. Thus the positions of such systems in a plot of flux vs.\ projected separation are ambiguous. For systems that are not detected, we know that neither the close pair nor the wide companion has substantial submillimeter flux; thus the system is reasonably represented by two points, one at each projected separation.\footnote{However, a counter argument to this is that a hierarchical quadruple system is quite different from just a wide binary, since this ``wide pair'' would not be expected to have a large flux if both close binaries cleared their disks. The only quadruple in our sample is 162819$-$2423 (ROX 43A/B), which is represented by three data points, one at each projected separation in the system. However, the closest pair in the system (the spectroscopic binary 162819$-$2423S) has a separation smaller than the 1 AU lower bound of our sample. Thus only two points for this system appear on the plots and in the statistical analysis. } For detected systems, the flux could arise from the close pair, the wide companion, or both. For these systems we have also plotted two points, one at each projected separation. For the pair with the larger projected separation, at least one member (whether that member is single or multiple) of the wide pair has detectable submillimeter flux. Thus, there is still detectable flux in the presence of a wide companion which should be represented by a point on the plot at the wider projected separation. For the pair with the closer projected separation, there {\em may\/} be detectable submillimeter flux from a close pair, and thus we adopt the conservative approach of putting a point on the plot to represent this possibility. Since we have argued in Paper I that close pairs have smaller fluxes on average than wide pairs, this method of assigning fluxes provides the most rigorous test of our hypothesis. Two triple systems, SR 24 and UZ Tau, have been treated differently because the wide pair in each has been resolved at millimeter wavelengths. SR 24 N and S are separated by 6\arcsec, and SR 24N is a close binary with a projected separation of 0\farcs2 (Simon et al.\ 1995). AM mapped the area around SR 24 using an ``on-the-fly'' mapping technique and a 12\arcsec\ (FWHM) beam, and they conclude that the detected 1300 \micron\ emission arises from an unresolved area around SR 24S, with an upper limit of 30 mJy on emission from SR 24N, the close binary. They give fluxes of 280 mJy for SR 24S and $<30$ mJy for SR 24N, though they note that the latter limit is uncertain due to possible contamination from SR 24S\null. This is consistent with the results of Reipurth et al.\ (1993), who found the total SR 24 flux in a 23\arcsec\ beam to be $259 \pm 14$ mJy at 1300 \micron. We adopt the flux values given by AM for our analysis. For UZ Tau, the wide pair has been resolved at 1300 \micron\ in interferometric observations using the Owens Valley Millimeter Array (Jensen, Koerner, \& Mathieu 1995, in preparation). UZ Tau W, the close binary, has a flux of $32 \pm 9$ mJy at 1300 \micron, while UZ Tau E has a flux of $137 \pm 28$ mJy. In our analysis we have assigned UZ Tau W its measured 1300 \micron\ flux and assigned the combined flux of both components to the wide E-W pair. Because of the availability of observations which resolve the wide pairs, these two systems are not marked as triples in Figures \ref{fig:sco-oph}--\ref{fig:clearing-model}. \subsection{The submillimeter flux distribution as a function of binary separation} Figure \ref{fig:sco-oph} shows the 1300 \micron\ flux of Sco-Oph binaries plotted as a function of projected binary separation. Filled symbols are detections and open symbols represent 3$\sigma$ upper limits. Squares show measurements at 1300 \micron, while triangles are 800 \micron\ measurements scaled to 1300 \micron\ as described above. To aid in determining which points may be confused by the presence of higher-order multiplicity, each point representing a pair in an unresolved higher-order multiple system is marked with a ``T'' to the right of the point. Different pairings from the same system can be associated by having the same flux measurement. (Note that in three instances the systems include a spectroscopic binary with a separation smaller than the limit of the figure.) All fluxes have been scaled to a distance of 140 pc, roughly the mean distance between the 125 pc distance to $\rho$ Oph and the 160 pc distance to Upper Sco. A dependence of submillimeter flux on projected binary separation is evident in Figure \ref{fig:sco-oph}. Binaries with projected separations $a_p$ less than roughly 100 AU tend to have lower fluxes than wider binaries. Among binaries with $a_p \ge 100$ AU, 7 of 15 are detected, while only 3 of 16 closer systems are detected, and two of these have separations very near 100 AU\null. More importantly, the close binaries which were not detected have flux upper limits substantially lower than most of the detections among the wider binaries. Clearly the submillimeter flux distribution differs between binaries with separations greater or less than roughly 100 AU, with the distribution extending to larger fluxes among the wider binaries. In order to quantify the dependence of submillimeter flux on binary separation, we applied statistical tests to compare the distributions of flux among different groups of binaries. We used the techniques of survival analysis which allow quantitative analysis of data which include upper limits. We used the software package ASURV Rev.\ 1.1 (LaValley, Isobe, \& Feigelson 1992) which implements the methods presented in Feigelson \& Nelson (1985). We divided the data into subsets based on projected separation and then compared pairs of these subsets using the Gehan, logrank, Peto-Peto, and Peto-Prentice two-sample tests. These tests yield probabilities that the distributions of flux in the two samples are drawn from the same parent distribution. The various tests are sensitive to different underlying flux distributions; since the true flux distribution for our sample is unknown, we have applied all of the tests and report the spread of the results. We have divided the sample into close and wide binaries using separations of both 50 AU and 100 AU as in Paper I\null. The results of the two-sample tests are given in Table~\ref{table:asurv_results}, expressed as percentage confidence levels that the flux distributions of the two samples are different. These results support the conclusion reached by visual examination of Figure \ref{fig:sco-oph}: at the 90--95\% confidence level binaries with projected separations less than or equal to 100 AU do not have the same submillimeter flux distribution as binaries with greater projected separations. When the sample is divided at 50 AU, the same result is found with similar confidence levels. Simon et al.\ (1992, 1995) found that young binaries as a group (i.e.\ of all separations) have lower submillimeter fluxes on average than young single stars. We have also compared the sample of ``single'' stars to the known binaries, but we have considered the close and wide binaries separately. The confidence levels at which the distributions of binaries and single stars differ are given in Table~\ref{table:asurv_results}. The close binaries have a different flux distribution from the single stars at confidence levels around 95\%. In contrast, the fluxes from the wide binaries and single stars are consistent with being drawn from the same distribution. We have applied the same two-sample tests discussed above to the PMS binaries and single stars in the Taurus-Auriga star-forming region. These results are also given in Table~\ref{table:asurv_results}, and the Taurus-Auriga binaries are plotted in Figure \ref{fig:tau-aur}. As expected, the results are the same as from similar previous analyses (Paper I; Osterloh \& Beckwith 1995): the submillimeter flux distribution of the close binaries is distinct from that of the wide binaries at high confidence levels. Thus, the same variation of submillimeter flux with binary separation is seen in two independent samples of PMS stars. Finally, we combine the Sco-Oph and Tau-Aur samples (Figure \ref{fig:combined-sample}), obtaining confidence levels greater than 99\% that the submillimeter flux distributions of close and wide binaries, and of close binaries and single stars, are distinct. Hence we conclude that at a statistically secure level the submillimeter flux distribution among binaries with separations between 1 AU and 50--100 AU is different than that of wider binaries or single stars. The sense of the difference is that submillimeter emission is weaker in close binaries. The binary separation at which the transition occurs is not well defined but appears to be roughly 50 AU to 100 AU\null. This separation is comparable to the radius typically derived for dust disks around PMS stars (e.g.\ Beckwith \& Sargent 1993). As pointed out by Osterloh \& Beckwith (1995), this transition separation strongly suggests that the emitting material does lie in disks in most cases. If the material were in extended envelopes (e.g.\ Terebey et al.\ 1993), binaries at this separation would not be expected to influence the submillimeter emission. The low submillimeter fluxes from close binaries are not due to age; the closer binaries have the same distribution of ages as the wider binaries or single stars. In addition, within the range of ages (roughly $10^5$--$10^7$ yr) in our sample, we find no dependence of submillimeter flux on age of the system for the wide binaries. \section{Interpreting the relationship between disk emission and binary separation} \subsection{Resonant clearing of gaps in disks} The presence of a stellar companion at a separation less than the disk radius must influence the spatial distribution of the disk material. Theoretical calculations and numerical simulations show that a binary embedded within a disk will rapidly clear a region on the size scale of the binary separation, thus isolating circumstellar and circumbinary disks (see e.g.\ Lin \& Papaloizou 1993 for a review). The extent of the region cleared depends on the details of the given system, specifically the binary mass ratio, orbital eccentricity, and disk viscosity. Artymowicz \& Lubow (1994) find that circumstellar disks will have outer radii of less than half the binary semi-major axis, and circumbinary disks will have inner radii of greater than roughly twice the binary semi-major axis. Disk clearing of this type creates a natural link between binary separation and the submillimeter fluxes from binary systems which is very similar to that observed. Assume that disks have a certain characteristic outer radius $R_d$. For the purposes of this discussion, this radius could be either a physical limit on the extent of disk material or an effective radius inside which most of the disk's submillimeter emission originates (even though there may be additional cold or low-density material outside it). In binary systems with separations greater than a few times $R_d$, each component of the system could have a circumstellar disk that is relatively undisturbed by its companion, and the submillimeter emission from such a system would be comparable to that from a single star with a circumstellar disk. In systems with separations much less than $R_d$ only a small hole at the center of the circumbinary disk will be cleared, leaving a circumbinary disk which could be similar in extent and emission to disks around single stars. Binaries with separations somewhat less than $R_d$, however, will clear regions whose extent represents a large fraction of the surface area of an undisturbed disk. This reduction in emitting surface area will result in reduced submillimeter fluxes, as observed. In order to investigate the effect of gap clearing on submillimeter flux in a more quantitative way, we have introduced gaps into a simple disk model. Following a standard approach in modeling infrared and submillimeter disk emission (e.g.\ Adams, Lada, \& Shu 1988, Adams, Emerson, \& Fuller 1990, BSCG), we use a geometrically thin disk with power-law temperature and surface-density distributions: \begin{equation}\label{eq:temperature} T(r) = T_1(r/{\rm (1\ AU)})^{-q}, \end{equation} \begin{equation} \Sigma(r) = \Sigma_0(r/r_0)^{-p}. \end{equation} The emission from such a disk is then given by \begin{equation}\label{eq:disk_emission} F_{\nu} = \frac{\cos\theta}{D^2} \int_{r_0}^{R_d} B_{\nu}[T(r)] (1 - e^{-\tau_\nu(r)})2\pi r \, dr, \end{equation} where $D$ is the distance from the observer to the system, $\theta$ is the angle between the line of sight and the normal to the disk plane, and $\tau_\nu$ is the line-of-sight optical depth through the disk, related to the mass opacity $\kappa_\nu$ by \begin{equation} \tau_\nu(r) = \frac{\kappa_\nu \Sigma(r)}{\cos \theta}. \end{equation} To simulate disk clearing by a binary companion, we made a simple modification to this disk model. For each model binary-disk system, we set the disk surface density (and thus the emission) equal to zero between radii $r_{in}$ and $r_{out}$, designating the inner and outer edges of the gap: \begin{equation}\label{eq:surface_density} \Sigma(r) = \cases{0,& if $r_{in} < r < r_{out}$;\cr \Sigma_0(r/r_0)^{-p},& otherwise.} \end{equation} These radii are determined by the parameters of a given binary system, especially the binary semi-major axis $a$. We considered two different cases. In the first, $r_{in} = 0.4a$ and $r_{out} = 1.8a$. This is the clearing expected for a binary with a circular orbit, and the minimum clearing expected for any binary system (Artymowicz \& Lubow 1994). In the second case, $r_{in} = 0.2 a$ and $r_{out} = 3a$. This level of clearing is expected for a binary with an orbital eccentricity of 0.4, roughly the mean eccentricity for pre--main-sequence\ and main-sequence binaries with known orbits (Mathieu 1994, Duquennoy \& Mayor 1991).\footnote{These clearing radii also depend on the disk viscosity and the binary mass ratio. The numbers quoted here are for a Reynolds number of $10^5$ and a binary mass ratio of 7:3. However, changing these values within reasonably expected bounds does not greatly affect the expected disk clearing. For a range of Reynolds numbers from $10^4$ to $10^6$ and eccentricity of 0.4, the inner gap edge ranges from $0.19a$ to $0.25a$ and the outer gap edge from $2.7a$ to $3.1a$ (Artymowicz \& Lubow 1994). We adopt the numbers above as representative of an ``average'' binary.} For a given set of disk parameters ($M_d$, $q$, $T_1$; $p = 1.5$, $R_d = 100$ AU) we calculated the expected submillimeter flux from binaries having a range of separations from 0.01 AU to 2000 AU, spanning the observed range of separations in our sample. We then calculated models with a range of inclination angles and averaged their fluxes based on the frequency of occurrence of a given value of $\theta$ in a random distribution of disk inclinations. The result for one particular disk model is shown in Figure \ref{fig:clearing-model}. The model is superimposed on the combined submillimeter data from the Sco-Oph and Tau-Aur regions, with all fluxes scaled to a distance of 140 pc. The solid line shows the expected submillimeter flux as a function of separation if gaps extend from $0.2 a$ to $3 a$ (eccentric orbits), and the dashed line shows the expected flux if gaps extend from $0.4 a$ to $1.8 a$ (circular orbits). As expected, the gap-clearing model shows the submillimeter fluxes of the closest and widest binaries to be unaffected by gap clearing, while binaries of intermediate separations have lower fluxes. For the particular model shown the reduction in flux occurs for binaries with separations between roughly 1 AU and 300 AU, with a minimum flux around 25 AU\null. Thus our simple model for gap clearing reproduces the qualitative features of the observed submillimeter flux dependence on binary separation. If the gaps are of the size expected for eccentric binary systems, the reduction of submillimeter flux due to gaps can be as much as a factor of 15. This is comparable to the difference between typical detected fluxes among wide binaries and the available upper limits for the close binaries. The model shown in Figure \ref{fig:clearing-model} provides a specific case for quantitative discussion. This model has $M_d = 0.05\ \hbox{${M}_{\sun}$}$, $q = 0.65$, and $T_1 = 125\ {\rm K}$ and was chosen to match typical flux levels among detected wide binaries. The minimum submillimeter flux in this particular model is a factor of nine less than the flux of the disk with no gap. The dashed line (circular binary) lies well above most of the observed flux upper limits between 1 and 100 AU and thus is not consistent with the observations. However, most binaries have eccentric orbits. The solid line ($e=0.4$) passes through the upper limits for binaries with separations around 25 AU and exceeds the typical upper limits for binaries of somewhat larger and smaller separations by a factor of two to three at most. {\em The essential conclusion to be drawn from this simple model is that the lack of detected submillimeter emission among the closer binaries does not necessarily imply that disk material is not present, nor that disk surface densities are much lower than found among wide binaries or single stars. The large reduction in emitting surface area due to gap clearing can in and of itself reduce flux to levels comparable to many of the present upper limits.} Thus, while it is clear from these submillimeter data that binary companions with separations less than 100 AU significantly influence the nature of associated disks, it remains an outstanding question to what extent circumstellar and circumbinary disks are present in these binaries. We return to these issues in Sections \ref{sec:cs-disks} and \ref{sec:cb-disks} Given the presence of a companion, the physical conditions of the disk(s) in our simple model are certainly oversimplified, and thus a detailed comparison of the model with observations is likely not merited. Nonetheless, we note that the model appears less successful for binaries with separations between 1 AU and 10 AU\null. In particular, the model flux is symmetric in $\log(a_p)$ around binary separations of a few tens of AU, while in fact submillimeter fluxes remain low at separations less than 10 AU\null. Unfortunately, the sample size in this separation range is small; the two-sample tests give probabilities of only 78--90\% that the submillimeter fluxes of binaries with $1 < a_p < 10$ AU are drawn from a different distribution than the fluxes of binaries with $a_p > 100$ AU\null. Furthermore, most of the binaries with $1 < a_p < 10$ AU were discovered via lunar occultation techniques, so that the measured separation is a projection against the lunar limb. As such, many of these binaries may have wider separations projected on the sky than those shown in Figure \ref{fig:clearing-model}. Thus, while there is a suggestion that there is a larger reduction in flux among the closest binaries than predicted by the model, the case is not strong with the present data. If the reduction in submillimeter flux does extend to separations as small as 1 AU, this would be difficult to reproduce with our simple annular gap model. We note that the flux predicted by the model for binaries with separations of less than 10 AU is largely from circumbinary disks. Indeed for binaries with separations of only a few AU the surface areas of circumstellar disks are too small to produce detectable submillimeter emission. Hence weak submillimeter emission from such binaries would suggest lower mass, temperature, and/or dust opacity of circumbinary material. Finally, we note that the most submillimeter-luminous binaries, such as GG Tau, T Tau, and AS 205, cannot be easily incorporated into this simple picture. GG Tau in particular has a projected separation of 40 AU and yet is the most luminous binary in the Tau-Aur and Sco-Oph regions. Furthermore, interferometric observations at millimeter wavelengths reveal a circumbinary disk with a radial extent of at least 800 AU (Dutrey, Guilloteau, \& Simon 1994). The millimeter emission also shows a central depression which Dutrey et al.\ attribute to a central cavity of radius 180 AU\null. They attribute this cavity to a clearing process similar to that invoked here for gaps, but clearly this system is substantially different from both our standard model for a 40 AU binary and from other binaries with projected separations of tens of AU\null. \subsubsection{Effect of disk parameters on the flux-separation relation} \label{sec:disk-params} The model shown in Figure \ref{fig:clearing-model} is only one realization of a large range of physical conditions found in disks. Therefore it is useful to establish whether other combinations of model parameters can better reproduce the observed variation of flux with separation. In Figure \ref{fig:4-clearing-models} we explore the influence of disk parameters on the distribution of submillimeter flux with binary separation. In panels a, b, and c, one of the disk parameters $T_1$, $q$, and $M_d$ is varied while the others are held fixed. In panel d, all three disk parameters are held fixed while the wavelength at which the flux is calculated is varied. One general trend that can be seen in these figures is that changes in disk properties which give lower submillimeter fluxes also tend to decrease the binary separation corresponding to the minimum flux. This is because the contribution of the hotter, more optically thick inner regions of the disk contribute a larger percentage of the total flux when the disk luminosity is low. For more luminous disks whose outer regions are hotter (through higher $T_1$ or lower $q$) or more optically thick (through higher $M_d$), the contribution of the outer disk to the total flux is significant. In such disks gap clearing in the outer regions has a larger effect on the total submillimeter flux. A discontinuous change in slope can be seen in some of the models in Figure \ref{fig:4-clearing-models} (as well as in Figure \ref{fig:clearing-model}). This is due to the assumption in the models that the disks have a sharp outer radius $R_d$, here taken to be 100 AU\null. The slope discontinuity occurs when the outer edge of the gap $r_{out}$ reaches the outer edge of the disk. Increasing $R_d$ would increase the separation that gives the minimum flux for a few of the models shown in Figure \ref{fig:4-clearing-models}, though most would be unchanged. Interestingly, the choice of 1300 \micron\ or 800 \micron\ as an observing wavelength makes little difference in the binary separation producing the minimum flux (Figure \ref{fig:4-clearing-models}d). This is due to the interplay of two opposing factors. Shorter-wavelength emission is more sensitive to disk material at higher temperatures, suggesting that 800 \micron\ emission would tend to probe gaps at smaller disk radii than 1300 \micron\ emission. However, for emission in the Rayleigh limit the tradeoff between increasing surface area and decreasing optical depth with radius tends to emphasize the emission from the region in the disk near $\tau_\nu = 1$ (BSCG). Because dust opacity increases with frequency, $\tau_\nu = 1$ lies at a larger radius for 800 \micron\ than for 1300 \micron. These two effects largely cancel each other and emission at both wavelengths is sensitive to roughly the same disk radii. However, 60 \micron\ flux is sensitive to smaller disk radii because much of the disk is too cold to produce appreciable 60 \micron\ emission. We draw two main conclusions from this exploration of model parameter space. First, under no circumstances does gap clearing at the level expected for binaries with circular orbits produce a reduction in flux that is comparable to that observed. Second, no combination of model parameters does significantly better than that in Figure \ref{fig:clearing-model} at reproducing the depth, breadth, and location of the reduction in submillimeter fluxes observed among the close binaries. \subsubsection{Estimates of disk masses}\label{sec:masscalc} Gap clearing may affect not only the distribution of submillimeter flux with separation, but also the disk mass derived for a given system based on its submillimeter flux. Submillimeter fluxes have commonly been used to infer disk masses by adopting a model for a continuous disk, deriving the disk temperature distribution from infrared data, and then adjusting the disk mass until the model flux agrees with the observed submillimeter flux at one or more wavelengths (Adams et al.\ 1990, BSCG, AM, Osterloh \& Beckwith 1995). However, if binaries clear gaps in their disks, the geometry of the emitting material is significantly different than that assumed in conventional disk modeling and may affect the mass calculation. Here we derive disk masses assuming the disk-clearing model discussed above and compare them to masses derived assuming continuous disks. Because disk clearing has little effect on disk mass estimates for wider binaries, we derive masses only for Sco-Oph binaries with $a_p < 300$ AU\null. The temperature distribution of the disk is determined by fitting the 10--100 \micron\ data for each system with a power law and determining $q$ and $T_1$ from the parameters of the fit (BSCG). Some of the binaries in our sample do not have sufficient infrared data available in the literature for this to be possible, and we could not derive masses for these systems. For the remaining systems, we took published values of $q$ and $T_1$ from AM for some sources and used infrared data from the literature to derive $q$ and $T_1$ for the others. Though emission at 10--100 \micron\ is likely to be dominated by the inner part of the circumstellar disk, we assume that the derived temperature distribution applies to the whole disk. To calculate the disk mass in the presence of a gap, we assume the emission from the disk is given by Eq.\ \ref{eq:disk_emission}, with the disk surface density given by Eq.\ \ref{eq:surface_density}. We calculate masses for three cases: no clearing, circular-orbit clearing ($r_{in} = 0.4a$, $r_{out} = 1.8a$), and eccentric-orbit clearing ($r_{in} = 0.2a$, $r_{out} = 3a$). For triple or quadruple systems, we choose the projected separation closest to 50 AU since this pair will most affect the derived disk mass. Following BSCG, we take $p = 1.5$, $\theta = 0$, $r_0 = 0.01$ AU, and $R_d = 100$ AU\null. Following Beckwith \& Sargent (1991), we use the opacity law $\kappa_\nu = 0.1\,(\lambda/250\ \micron)^{-\beta}$, with $\beta = 1$. We then numerically integrate Eq.\ \ref{eq:disk_emission} and vary $\Sigma_0$ until the calculated flux matches the observed flux. These masses are given in Table \ref{table:disk_masses}. The table shows that gap clearing does not greatly affect the derived mass for most systems. We further explored this affect using an ensemble of model binary systems with a range of disk properties. For gaps from $0.4 a$ to $1.8 a$, the derived disk mass typically differs by a factor of two or less from the mass calculated assuming a continuous disk; for gaps from $0.2 a$ to $3 a$, the difference is typically a factor of three or less. For some combinations of disk properties and gap locations, the variations are somewhat greater. Gap clearing can either increase or decrease the mass derived from a given flux, depending on whether the area cleared is efficient or inefficient at radiating submillimeter flux compared to the rest of the disk. Thus, gap clearing introduces an additional uncertainty into disk mass calculations for close binary systems. \subsection{Circumstellar disks in binary environments}\label{sec:cs-disks} While gap clearing can plausibly explain the reduction of submillimeter flux from binaries with separations of less than 50-100 AU, the upper limits on fluxes from such binaries can be equally well explained if the physical conditions (such as surface densities or temperatures) of disks in closer binaries differ from those in wide binaries. Indeed, one straightforward interpretation of the lack of submillimeter emission is the complete absence of disk material. However, 2 \micron\ and 10 \micron\ excesses indicate that circumstellar disks are present in at least 50\% of PMS binaries (Mathieu 1994). Similarly, Simon et al.\ (1995) find the same binary frequency in systems with and without circumstellar disks based on $K - L$ colors, and Simon \& Prato (1995) find the same frequency of $K - N$ color excesses for single stars as for binaries. But observations at these wavelengths only sample material very near stellar surfaces. In addition the high dust opacities at these wavelengths typically do not permit derivation of surface densities. Longer wavelength IRAS measurements also support a high frequency of circumstellar disks. Mid-infrared (e.g.\ 60 \micron) flux originates in the inner regions of disks (typically $\le 10$ AU); thus for most of the binaries in our sample 60 \micron\ flux originates in circumstellar disks. Figure \ref{fig:iras} shows IRAS 60 \micron\ flux plotted as a function of projected separation for the Tau-Aur and Sco-Oph binaries. IRAS fluxes or upper limits were taken from Weaver \& Jones (1992), Strom et al.\ (1989), Clark (1991), Hartmann et al.\ (1991), Wilking, Lada, \& Young (1989), and the IRAS Point Source Catalog. Filled symbols represent detections and open symbols are upper limits. All fluxes are scaled to a distance of 140 pc. Fourteen systems out of 85 do not have IRAS measurements in any of the above references and are not shown here. These systems are approximately equally divided between wide and close binaries. Figure \ref{fig:iras} has two notable features. First, the fraction of binaries detected at 60 \micron\ is much higher than at submillimeter wavelengths. If 60 \micron\ flux is taken to originate in disks, then at least one circumstellar disk is present in most PMS binaries. As such, the low level of submillimeter emission from the close binaries is not the result of a total absence of disk material. Second, there is no marked dependence of 60 \micron\ flux on binary separation akin to that seen at submillimeter wavelengths. The same two-sample tests performed on the submillimeter data show no difference in the 60 \micron\ flux distributions of the close and wide binaries divided at 10, 50, or 100 AU\null. These 60 \micron\ flux measurements can provide meaningful constraints on the surface-density normalizations of circumstellar disks. Because dust opacity is much higher at 60 \micron\ than at submillimeter wavelengths, disks remain optically thick at 60 \micron\ to much lower surface densities than in the submillimeter. An explicit assumption of previous analyses of disk masses is that the disks are optically thick at 60 \micron\ and thus reflect the disk temperature distributions. However, given only an upper limit on submillimeter flux and thus circumstellar disk surface density, this assumption need not hold. For a power-law radial temperature distribution one signature of optically thin 60 \micron\ emission would be a steepening of the spectral slope between 12 \micron\ and 60 \micron. In fact, several binaries show such steepening at a formally significant level given the quoted uncertainties on the IRAS fluxes. Thus some circumstellar disks may be partially optically thin at 60 \micron. However, given that other binaries also show deviations from single spectral slopes in other senses (e.g., high 12 \micron, 25 \micron, or 60 \micron\ fluxes at formally significant levels), we do not feel that in any given case the conclusion of optical thinness is secure. Rather we choose to use the binaries with steepening spectral slopes to derive lower limits on masses and surface densities of circumstellar disks. Specifically we have chosen three Tau-Aur binaries with $a_p < 50$ AU and for which the slope of $\lambda F_\lambda$ from 25 to 60 \micron\ is steeper than the slope from 12 to 25 \micron: DF Tau, FO Tau, and CZ Tau. These binaries have the highest quality IRAS flux measurements and no additional companions.\footnote{CZ Tau lies 30\arcsec\ from DD Tau, but Weaver \& Jones (1992) list CZ Tau as a ``better positional fit'' to the IRAS source.} The IRAS fluxes of these three binaries lie in the middle of the range of detected fluxes, and thus we take these systems to be representative examples of young binary systems. For each binary we have derived disk masses using the method described in \S \ref{sec:masscalc}, except that the disk temperature was derived from only the IRAS 12 and 25 \micron\ fluxes and the disk radius was taken to be half the projected binary separation. The disk surface density was then varied so that the model flux matched the observed 60 \micron\ flux. We used the dust opacity law given by Adams et al.\ (1988) since Beckwith \& Sargent (1991) do not give opacities for IRAS wavelengths. The circumstellar disk masses derived for these binaries range from $5 \times 10^{-6}$ \hbox{${M}_{\sun}$}\ to $7 \times 10^{-5}$ \hbox{${M}_{\sun}$}. These numbers cannot be directly compared with masses derived for disks around single stars because of the smaller disk radii used here. A more significant comparison is the disk surface-density normalization ($\Sigma_0$; Eq.\ \ref{eq:surface_density}). The derived surface-density normalizations for these circumstellar disks are roughly two orders of magnitude smaller than those for disks around single stars detected at submillimeter wavelengths. {\em The essential conclusion to be drawn from this analysis is that the mechanism reducing the submillimeter flux from binaries with $1 < a_p < $ 50--100 AU does not entirely destroy circumstellar disks or inhibit their formation.} We stress that these mass and surface density estimates are best considered as lower limits. While the steepening spectral slopes of these binaries may be indicative of partial optical thinness at 60 \micron\ as presumed in these calculations, we cannot rule out fluctuations in the IRAS fluxes larger than the formal errors. At the same time, we have shown that in the presence of gaps the submillimeter flux upper limits do not require that the circumstellar surface densities be significantly lower than found in disks around single stars. Thus, the available data require circumstellar disk surface densities in close binaries to be in a range from 1\% to the same as typical disks around single stars. These limits constrain the degree of depletion of circumstellar disks by the various processes discussed in the Introduction. The lack of dependence of 60 \micron\ emission on binary separation is not inconsistent with the gap model. The specific model shown in Figure \ref{fig:4-clearing-models}d predicts a factor of $\sim 5$ decrease in 60 \micron\ flux for binaries with separations between 1 AU and 10 AU\null. Given the small number of systems in this separation range, such a change would be undetectable with current data. In addition, as noted previously most binaries at these small separations were discovered via lunar occultation, so that their true separations may be substantially underestimated. Finally, it is plausible that the temperature distributions of disks in close binary environments would differ from those in wide binaries or around single stars, also influencing submillimeter and infrared emission. We used the $T_1$ and $q$ values (see Eq.\ \ref{eq:temperature}) for the Sco-Oph and Tau-Aur binaries to investigate whether the temperature parameters vary systematically with separation. We find no evidence for a dependence of disk temperature on binary separation. These parameters only provide information about the inner 10 AU or so of the disk. The temperature distribution of the disks elsewhere---and in particular at radii comparable to most binary separations---is largely unconstrained by current observations. \subsection{Circumbinary disks} \label{sec:cb-disks} Because of their low temperatures, circumbinary disks in all but the closest binaries emit predominantly at far-infrared and longer wavelengths. However, since substantial emission at these wavelengths can arise from a large range of radii within a disk, securely associating unresolved flux measurements with circumbinary disks can be problematic. On the other hand, upper limits on submillimeter emission unambiguously place upper limits on circumbinary disk masses. For optically-thin emission at $\lambda = 1300$ \micron\ with $\kappa_\nu = 0.02$ cm$^{2}$ g$^{-1}$, the circumbinary mass $M_{cb}$ that produces a given flux $F_\nu$ is \begin{equation}\label{eq:cb-mass} M_{cb} = 2.6 \times 10^{-4}\, \hbox{${M}_{\sun}$}\ (F_\nu/{\rm mJy})\ (e^{(11.1\ {\rm K}/T)} - 1) \end{equation} assuming a distance of 140 pc and emission in the Rayleigh limit. Taking $T=15$ K as a minimum dust temperature and $F_\nu < 15$ mJy as a typical flux upper limit for binaries with separations of less than $\sim 100$ AU, we find $M_{cb} < 4.3 \times 10^{-3}$ \hbox{${M}_{\sun}$}. This upper limit is conservative given that some emission may derive from a circumstellar disk. On the other hand if the circumbinary material were optically thick, the derived mass limit would be higher. {\em Thus we conclude that binaries with projected separations between a few AU and $\sim 100$ AU typically do not have circumbinary disks with masses greater than 0.005 \hbox{${M}_{\sun}$}.} There is one notable counterexample, however. As discussed above, a ring-like circumbinary disk around the 40 AU binary GG Tau has been resolved at 3 mm (Dutrey et al.\ 1994). The mass derived for this ring from the continuum emission is 0.13 \hbox{${M}_{\sun}$}. GG Tau is unusual in being the most luminous submillimeter source in our sample (Figure \ref{fig:combined-sample}) and one of the younger binaries in Tau-Aur. Most PMS binaries of similar projected separation do not have similarly massive or luminous circumbinary disks, suggesting the possibility that GG Tau represents a brief phase in early circumbinary disk evolution. Circumbinary disks have also been found around several very close binaries. For example, in HP Tau (one of the closest binaries in our sample with detected 1300 \micron\ emission), a simple calculation shows that a circumbinary disk must be present. If the binary orbit were filled with optically-thick material emitting at the stellar temperature, the resulting submillimeter flux would be comparable to that observed; however, the optical flux of such a quantity of hot material would be $\sim 10^4$ times the observed flux. Thus, the submillimeter-emitting region in HP Tau must be substantially larger than a few AU in radius and therefore must lie primarily outside the binary orbit. A similar calculation requires the presence of circumbinary disks around the PMS spectroscopic binaries GW Ori (Mathieu et al.\ 1995), AK Sco and V4046 Sgr (detected in this work), and DQ Tau (Mathieu 1995). To conclude, our flux limits place upper limits on circumbinary disk masses of 0.005 \hbox{${M}_{\sun}$}\ for most binaries with separations of less than roughly 100 AU\null. Indeed, it remains possible that most binaries do not have circumbinary disks. The present submillimeter flux detections and upper limits for almost all binaries with separations greater than a few AU are entirely consistent with emission from only circumstellar disks. \section{Conclusion} We have made sensitive 800 \micron\ continuum observations of most (25) pre--main-sequence\ binaries with projected separations $a_p \lesssim 150$ AU in the Scorpius-Ophiuchus star-forming region, and we have supplemented these data with previous 1300 \micron\ continuum observations to obtain a sample of 30\ systems. We have also created a similar database from the literature for 55\ systems in Taurus-Auriga. We have used these data to study the nature of disks in young binary environments, and we find: 1) Submillimeter fluxes from binaries with $1 < a_p <$ 50--100 AU are lower on average than from wider binaries or single stars, whereas the flux distributions of wide binaries and single stars are indistinguishable. This dependence of submillimeter flux on binary separation is seen independently in the Sco-Oph and Tau-Aur samples. When the samples are combined the effect is found at greater than the 99\% confidence level. The transition separation of 50--100 AU is similar to typically derived radii for dust disks around young stars, strongly suggesting that the reduction in submillimeter emission among closer binaries is due to the influence of the companions on disks. 2) The reduction in submillimeter flux from binaries with $1 < a_p <$ 50--100 AU can plausibly be attributed to gaps cleared in disks by binaries with eccentric orbits. As such, the present upper limits permit but do not require a large reduction in the surface densities of disk material outside such gaps compared to surface densities of disks among wide binaries or single stars. 3) Most of the binaries in our sample were detected at 60 \micron\ by IRAS, indicating that each of these binaries has at least one circumstellar disk. Presuming the disks to be optically thick at 60 \micron, the flux measurements place lower limits of roughly $10^{-5}$ \hbox{${M}_{\sun}$}\ on circumstellar disk masses. This lower limit corresponds to circumstellar disk surface densities no more than two orders of magnitude smaller than surface densities of most disks detected at submillimeter wavelengths. 4) We place upper limits of 0.005 \hbox{${M}_{\sun}$}\ on circumbinary disk masses around most binaries with projected separations between 1 AU and 100 AU\null. Circumbinary disks as massive as that found around GG Tau are rare for these binaries. However, submillimeter detections of binaries with separations less than a few AU show that massive circumbinary disks can exist around the closest binaries. The present body of data is consistent with the following picture for disks in PMS binary environments. Binaries with semimajor axes of a few hundred AU or greater have circumstellar disks with properties very similar to those of disks found around single stars. Binaries with semimajor axes between a few AU and 50--100 AU also typically have at least one circumstellar disk each. The binaries truncate these disks, thus limiting their submillimeter emission. Their surface densities remain uncertain within a range of 0.01 to 1 times the surface densities of disks around single stars. Binaries in this separation range typically do not have massive ($M > 0.005$ \hbox{${M}_{\sun}$}) circumbinary disks. Binaries with semimajor axes of less than a few AU can have massive circumbinary disks. Such binaries truncate the inner edges of circumbinary disks at radii of only a few AU or less, leaving most of the disk undisturbed on dynamical timescales. Notably absent from this morphological picture is any discussion of the detailed physical conditions of the disks, such as temperature, surface-density, and opacity distributions in both circumstellar and circumbinary disks. In addition to clearing gaps, companions are expected to influence these disk properties (e.g.\ Syer \& Clarke 1995). At the same time, the present submillimeter flux upper limits can only just be explained by our model, even assuming relatively large gaps driven by eccentric binaries. We anticipate that disks in young binary environments are substantially more complex than the simple disk model employed here. Finally, we note that binaries are the primary product of star formation, and thus the frequency of other planetary systems is intimately linked to the issues raised here. For binaries with separations of less than 100 AU, present upper limits on submillimeter flux imply upper limits on total disk mass of a few times $10^{-3}$ \hbox{${M}_{\sun}$}. This is roughly an order of magnitude smaller than typical estimates for the minimum mass of the early solar nebula (e.g.\ Weidenschilling 1977). Such a disk could conceivably form terrestrial planets, but it is unlikely to form gaseous giant planets. However, these mass estimates are insensitive to material in grains or planetesimals larger than a few mm in size. Thus, it remains possible that close binaries can form planets, but it would appear that sufficient disk material to form planetary systems like our own is most likely to be found in wide binaries or single stars. \acknowledgements{ We thank the referee, Steve Beckwith, for useful comments which improved the presentation of this paper. We are grateful to the staff of the JCMT for their knowledgeable support and good humor. ELNJ gratefully acknowledges the support of a Grant-in-Aid of Research from the National Academy of Sciences through Sigma Xi, as well as funding from the National Space Grant College and Fellowship Program and the Wisconsin Space Grant Consortium. RDM appreciates funding from the Presidential Young Investigator program, a Guggenheim Fellowship, the Morrison Fund of Lick Observatory, and the Wisconsin Alumni Research Fund. GAF acknowledges the support of an NRAO Jansky Fellowship. This research has made use of the Simbad database, operated at CDS, Strasbourg, France. }
\section{Introduction} \noindent \noindent Effective actions of gauge theories are space--time integrals over gauge and Lorentz invariant expressions. From the mathematical point of view, they are, up to some factors, functional traces of heat kernel coefficients, known as Schwinger--DeWitt,\cite{SD} Gilkey--Seeley,\cite{GS} or Hadamard coefficients.\cite{H} In flat space-time, these coefficients are polynomials constructed from a matrix potential and from the gauge field strength tensor by multiplication, gauge covariant differentiation, and contraction of Lorentz indices. Due to Bianchi identities and the product rule for covariant derivatives, the form of the coefficients is not unique. Furthermore, the physically interesting functional trace of the coefficients allows cyclic exchanges of matrix factors and integration by parts. New methods of computing effective actions, such as the string--inspired world line path integral formalism,\cite{worldline formalism,fliegner} but also the implementation of established calculation algorithms on computers\cite{lanyov} enable the extension of known results to higher order in the inverse mass expansion. To manage the corresponding increasing number of terms and to compare results of different methods,\cite{methods,results} a standard basis of invariants is needed, in terms of which all results can be expressed. An algorithm should be provided to convert a Lorentz scalar given in a non--standard form into terms of the basis. For gravitational invariants, constructed from the Riemann and the metric tensor, such normal forms were presented up to order eight in the mass dimension by Fulling et al.\cite{Fulling et al.} In the general case with matter, gauge fields, and gravity, basis sets of non--local invariants up to third order in the curvature were constructed. They are used in the expansion of effective actions in terms of Barvinsky--Vilkovisky form factors.\cite{BV} This contribution analyzes the formal structure of invariant monomials in non--Abelian gauge theories with matter in flat space--time. Step by step, the operations applicable to invariants are used to convert them into a fixed form. Thus, a basis of invariants is specified, and simultaneously, a procedure to expand an arbitrary given Lorentz invariant expression in terms of the basis is obtained. The proof of the basis property of the specified set of invariants will be published elsewhere.\cite{proof} \section{Notations} \label{graph.not} \noindent Notations are introduced on the basis of a concrete example. Let us consider a gauged scalar field theory described by the massive complex field $\phi^a$ and the Hermitian matrix valued gauge field $A_\mu^{ab}$. The gauge covariant derivative in the fundamental representation is ${\cal D}_\mu^{ab}=\delta^{ab}\partial_\mu-\i A_\mu^{ab}$. The coupling constant is contained in the gauge field. Integrating the quantum fluctuations of the field $\phi^a$ in the given backgrounds $\varphi^a$ and $A_\mu^a$, we obtain, in a first approximation, the one--loop effective action $\Gamma^{(1)}[\varphi,A]$ which can be expanded in gauge invariant terms\cite{methods} \vspace*{-0.2cm} \begin{equation}\label{generic result} \Gamma^{(1)}[\varphi,A]={\rm\;Tr}\ln\left(-{\cal D}^2+V+m^2\right)= \int\d^dx\sum_i\frac{C_i}{m^{\mu_i-d}}{\rm\;tr}\left(I_i\left(F,V\right)\right). \vspace*{-0.3cm} \end{equation} $V$ is a matrix potential originating from the matter fields. The $C_i$ are complex numbers and $I_i\left(F,V\right)$ matrix valued Lorentz scalars composed of the potential $V$, the field strength tensor \begin{math} F_{\mu\nu}^{ab}=\i\comm{{\cal D}_\mu}{{\cal D}_\nu}^{ab}= \partial_\mu A_\nu^{ab}-\partial_\nu A_\mu^{ab}- \i\comm{A_\mu}{A_\nu}^{ab}, \end{math} and the gauge covariant derivative in the adjoint representation \begin{math} D_\mu =\comm{{\cal D}_\mu}{.}=\partial_\mu -\i \comm{A_\mu}{.}. \end{math} $D_\mu$ acts on the matrix potential and on the field strength tensor. $d$ is the dimension of space--time. $\mu_i$ is the mass dimension of the scalar $I_i(F,V)$ according to the mass dimensions of its constituents $[V]=2$, $\left[F_{\mu\nu}\right]=2$, and $\left[{\cal D}_\mu\right]=\left[D_\mu\right]=1$. The form (\ref{generic result}) is not unique due to several equalities, namely the product rule for covariant derivatives, integration by parts, cyclic permutations, the Bianchi identity, the antisymmetry of the field strength tensor, and the exchange of derivatives: \begin{displaymath} \refstepcounter{equation} \label{manipulations} \begin{array}{c@{\quad}cr} D_\mu(XY)=D_\mu XY+XD_\mu Y,& \int\d x{\rm\;tr}\left(D_\mu X_\mu Y\right) =-\int\d x{\rm\;tr}\left(X_\mu D_\mu Y\right)\!,& \makebox[0.8cm][r]{(\theequation\rm a,b)\hspace{-0.2cm}}\\ {\rm\;tr}(XY\ldots Z)={\rm\;tr}(Y\ldots ZX), & D_\mu F_{\kappa\lambda}=D_\kappa F_{\mu\lambda}+D_\lambda F_{\kappa\mu},& \makebox[0.45cm][r]{(\theequation\rm c,d)\hspace{-0.2cm}}\\ F_{\mu\nu}=-F_{\nu\mu},& D_\mu D_\nu X=D_\nu D_\mu X-\i\comm{F_{\mu\nu}}{X}.& \makebox[0.45cm][r]{(\theequation\rm e,f)\hspace{-0.2cm}} \end{array} \end{displaymath} Let us call a $V$, an $F$, or covariant derivatives of them a {\em simple factor\/}, i.e. \vspace*{-0.1cm} \begin{equation}\label{factors} (\mbox{simple factor})\in\{ V,\; F_{\kappa\lambda},\; D_{\mu_1}D_{\mu_2}\ldots D_{\mu_n}V,\; D_{\mu_1}D_{\mu_2}\ldots D_{\mu_n}F_{\kappa\lambda}\}.\vspace*{-0.1cm} \end{equation} Simple factors containing the matrix potential are called {\em $V$--factors}, the others {\em $F$--factors}. With the product rule (\ref{manipulations}a), expression (\ref{generic result}) can be converted into a form where the invariants $I_i(F,V)$ are monomials, i.e. products of simple factors. Subsequently, the invariants are supposed to have this form. If the gauge group representation is unitary, the additional symmetries \vspace*{-0.15cm} \begin{equation}\label{symmetries} V^\dagger=V,\quad A_\mu^\dagger=A_\mu,\quad F_{\mu\nu}^\dagger=F_{\mu\nu},\quad \left(D_\mu X\right)^\dagger=D_\mu X\quad\mbox{if}\quad X^\dagger=X\vspace*{-0.15cm} \end{equation} hold. Consequently, simple factors are Hermitian. For simple factors $X$, $Y$, and $Z$ this leads to \vspace*{-0.4cm} \begin{equation}\label{mirror transform} \overline{{\rm\;tr}\left(XYZ\ldots\right)}= {\rm\;tr}\left(\ldots Z^\dagger Y^\dagger X^\dagger\right)= {\rm\;tr}\left(\ldots ZYX\right).\vspace*{-0.15cm} \end{equation} Thus, an invariant monomial can be expressed by the complex conjugate of its mirror image with identical factors, but in inverted order. Therefore we call eq.~(\ref{mirror transform}) a mirror transformation. In general, a monomial and its complex conjugate are independent of each other, so that operation (\ref{mirror transform}) cannot be used to reduce the number of terms in eq.~(\ref{generic result}). However, Lagrangians are real. Hence, in an appropriate basis, an arbitrary invariant monomial $I(F,V)$ and its mirror image have complex conjugate coefficients $C$ and $\bar{C}$ so that they add to $2\Re{\rm e}\left(C\cdot I(F,V)\right)$. An\-other exception occurs for real $\phi^a$ and imaginary $A_\mu^{ab}$.\footnote{This is the case for real orthogonal representations of the gauge group. Then $\i A_\mu^{ab}$ is real and antisymmetric in $a$ and $b$.} Then $V$--factors are real and $F$--factors imaginary. In this case, monomials and their complex conjugates are not independent of each other and eq.~(\ref{mirror transform}) reduces the number of terms in eq.~(\ref{generic result}) indeed. \section{The Basis} \label{construction} \vspace*{-0.4cm} \subsection{The reduction algorithm} \noindent We start from an arbitrary Lorentz invariant given in the form (\ref{generic result}). The product rule must be used whenever derivatives of products are encountered. This may happen at each stage of the algorithm. The manipulations (\ref{manipulations}b--f, \ref{mirror transform}) must be applied in the sequence of the following sub-subsections to obtain a standard result. The rules given there do not entirely fix all details of the algorithm. Therefore, the algorithm can be executed in different ways, but the results will be expressed by the same basis of invariants and, hence, will be identical. The procedure will require exchanges of derivatives by eq.~(\ref{manipulations}f). Since thereby additional invariants with more $F$--factors and fewer derivatives are produced, the algorithm starts with the invariants with the most $F$--factors and descends to invariants with fewer and fewer $F$--factors. \subsubsection{Integration by parts} \noindent The indices in a Lorentz invariant monomial can be contracted between different factors and within the same factor. We call the latter self--contractions. They always include a co\-var\-i\-ant derivative. Therefore, {\it we apply integration by parts to covariant derivatives in self--contractions.} Thereby all self--contractions are eliminated. \subsubsection{The Bianchi identity} \noindent The Bianchi identity (\ref{manipulations}d) exchanges the index of one derivative with the indices of $F_{\mu\nu}$ within an $F$--factor. All other factors remain unchanged. Therefore, we need a prescription that specifies the derivatives which are candidates for applying the Bianchi identity in the $F$--factor under consideration. Let us consider the example \vspace*{-0.3cm} \begin{equation}\label{example sectors} {\rm\;tr}( \underbrace{ \stackrel{\mbox{L}}{D_\mu} \stackrel{\mbox{R}}{D_\nu} \stackrel{\mbox{M}}{D_\rho} D_\sigma F_{\kappa\lambda}}_{\makebox[0cm]{\parbox{2.1cm}{\centering factor under\\consideration}}} \underbrace{\ldots X_\nu'\ldots}_{\parbox{1.5cm}{\centering right sector}} Y_{\sigma\kappa} \underbrace{\ldots X_\rho''\ldots}_{\parbox{1.5cm} {\centering middle sector}}Z_\lambda \underbrace{\rule{0cm}{1.85ex} \ldots X_\mu\ldots}_{\makebox[0cm][l]{\parbox{1.5cm}{\centering left sector}}})\quad.\vspace*{-0.1cm} \end{equation} The indices of $F_{\kappa\lambda}$ are contracted with the factors $Y_{\sigma\kappa}$ and $Z_\lambda$, which divide the remaining factors into three, possibly empty, sectors. We call them ``right sector'', ``middle sector'', and ``left sector'', as indicated, because, due to cyclic invariance (\ref{manipulations}c), the ``left sector'' is connected with the left--hand side of the factor under consideration. The derivatives of the factor under consideration are called left (``L''), right (``R''), and middle (``M'') corresponding to the sector they are contracted with. Not all derivatives are left, right, or middle (e.g.~$D_\sigma$). The Bianchi identity (\ref{manipulations}d) mixes all three kinds of derivatives. Therefore it can be used to eliminate one kind of index in all factors of all monomials. Since the middle sector is invariant under the mirror transformation (left and right sectors are interchanged), {\it we apply the Bianchi identity to middle derivatives.} Each such application of the Bianchi identity reduces the number of factors in the corresponding middle sector. Thus, after finitely many steps, all middle derivatives are eliminated. Finally, we convert multiple contractions between factors into a standard form by \begin{eqnarray}\label{multiple1}\textstyle \ldots F_{\mu\nu}\ldots D_\mu D_\nu X\ldots&\Rightarrow&\textstyle -\frac{\i}{2}\ldots F_{\mu\nu}\ldots\comm{F_{\mu\nu}}{X}\ldots\\ \label{multiple2}\textstyle \ldots F_{\mu\nu}\ldots D_\mu F_{\nu\kappa}\ldots&\Rightarrow&\textstyle \frac{1}{2}\ldots F_{\mu\nu}\ldots D_\kappa F_{\nu\mu}\ldots\\ \label{multiple3} \ldots D_\mu F_{\nu\kappa}\ldots D_\nu F_{\mu\lambda}\ldots&\Rightarrow& \ldots D_\mu F_{\nu\kappa}\ldots D_\mu F_{\nu\lambda}\ldots+\nonumber\\ &&\textstyle\hspace{1cm}+ \frac{1}{2}\ldots D_\kappa F_{\nu\mu}\ldots D_\lambda F_{\mu\nu}\ldots\quad. \end{eqnarray} The first equality uses the antisymmetry (\ref{manipulations}e) of the field strength tensor and the commutation rule (\ref{manipulations}f). The second transformation relies on the antisymmetry (\ref{manipulations}e) together with the Bianchi identity (\ref{manipulations}d). The third rule results by applying the Bianchi identity (\ref{manipulations}d) to one of the factors and subsequently using eq.~(\ref{multiple2}). \subsubsection{The arrangement of factors} \noindent Cyclic factor permutations (\ref{manipulations}c) and, possibly, mirror transformations (\ref{mirror transform}) can be used to identify invariants. Applying eqs. (\ref{manipulations}c) and (\ref{mirror transform}) in all possible ways to a given invariant monomial, we obtain a class of equivalent invariants. {\em We pick a representative of each equivalence class.} This may be done by introducing an ordering relation in the equivalence classes Then we pick the smallest (or greatest) invariant of each equivalence class as the representative. \subsubsection{The arrangement of indices} \label{index convention} \noindent Derivatives and indices of $F$'s can be exchanged by means of eqs.~(\ref{manipulations}f) and (\ref{manipulations}e) in all factors of all invariant monomials. Let us consider a certain factor within an invariant. It can be shifted completely to the left--hand side by eq.~(\ref{manipulations}c), as a result of which the achieved arrangement of factors is temporarily destroyed\footnote{The arrangement of the factors has to be restored after reordering the indices and is, in the end, not affected by this procedure.} \hspace{0.1cm} (cf.\ example~(\ref{example sectors})). {\it After this operation, we rearrange the derivatives and/or indices of the $F$ (if present) in the considered factor according to the contracted counter indices.\/} In example~(\ref{example sectors}) $Y_{\sigma\kappa}$ is located left of $Z_\lambda$. Thus the indices of $F_{\kappa\lambda}$ have the correct order. The locations of $X_\nu'$, $Y_{\sigma\kappa}$, $X_\rho''$, and $X_\mu$ define the correct order of the derivatives to be $D_\nu D_\sigma D_\rho D_\mu$. Since the mirror transformation inverts the ordering of the factors, it has to be applied {\it before\/} rearranging the indices. Cyclic factor permutations and the arrangement of indices do not interfere with each other. \subsection{The defining properties of the basis} \noindent Pursuing the above algorithm, we state the following properties of basis invariants:\vspace*{-0.2cm} \begin{itemlist} \item The invariants are products of simple factors. \item Indices are contracted only between different factors of an invariant monomial. \item There are no ``middle'' derivatives. \item In multiple contractions between factors, derivatives are contracted with de\-riv\-a\-tives and indices of $F$'s with indices of $F$'s (cf.\ eqs.~(\ref{multiple1} -- \ref{multiple3})) except for contractions of an index of an $F$ with a derivative where the other index of the $F$ is contracted with a third factor. \item The order of derivatives and of indices of the $F$'s is as described in sub-subsection 3.1.4.\vspace*{-0.2cm} \end{itemlist} These properties allow to count the basis invariants of a certain mass dimension. Up to mass dimension 16, this was performed by a C language program (table \ref{number of invariants}). Results of higher dimension or divided by the number of $F$'s are available. \begin{table}[htbp] \centering \tcaption{The number of basis invariants {\bf with} and {\it without} the mirror transformation. $v$ is the number of occurrences of the matrix potential $V$ in the invariants.} \label{number of invariants} \vspace*{1ex}\footnotesize\baselineskip=10pt \raggedright \begin{tabular}{|c|c||rr||rr|rr|rr|} \hline &Mass dim.&\multicolumn{2}{c||}{Total}& \multicolumn{2}{c|}{$v=0$}&\multicolumn{2}{c|}{1}& \multicolumn{2}{c|}{2}\\ \hline 1&2&\bf 1&\it 1&\bf 0&\it 0&\bf 1&\it 1&&\\ 2&4&\bf 2&\it 2&\bf 1&\it 1&\bf 0&\it 0&\bf 1&\it 1\\ 3&6&\bf 5&\it 5&\bf 2&\it 2&\bf 1&\it 1&\bf 1&\it 1\\ 4&8&\bf 17&\it 18&\bf 7&\it 7&\bf 4&\it 5&\bf 4&\it 4\\ 5&10&\bf 79&\it 105&\bf 29&\it 36&\bf 24&\it 36&\bf 17&\it 23\\ 6&12&\bf 554&\it 902&\bf 196&\it 300&\bf 184&\it 329&\bf 119&\it 191\\ 7&14&\bf 5283&\it 9749&\bf 1788&\it 3218&\bf 1911&\it 3655& \bf 1096&\it 2020\\ 8&16&\bf 65346&\it 127072&\bf 21994&\it 42335&\bf 24252&\it 47844& \bf 13333&\it 25861\\ \hline \end{tabular}\\[1ex] Table \ref{number of invariants}. (Continued)\\ \begin{tabular}{|c|c||rr|rr|rr|rr|rr|rr|} \hline &Mass dim.&\multicolumn{2}{c|}{$v=3$}&\multicolumn{2}{c|}{4}& \multicolumn{2}{c|}{5}&\multicolumn{2}{c|}{6}& \multicolumn{2}{c|}{7}&\multicolumn{2}{c|}{8}\\ \hline 1&2&&&&&&&&&&&&\\ 2&4&&&&&&&&&&&&\\ 3&6&\bf 1&\it 1&&&&&&&&&&\\ 4&8&\bf 1&\it 1&\bf 1&\it 1&&&&&&&&\\ 5&10&\bf 6&\it 7&\bf 2&\it 2&\bf 1&\it 1&&&&&&\\ 6&12&\bf 39&\it 63&\bf 13&\it 16&\bf 2&\it 2&\bf 1&\it 1&&&&\\ 7&14&\bf 370&\it 670&\bf 96&\it 158&\bf 18&\it 24&\bf 3&\it 3& \bf 1&\it 1&&\\ 8&16&\bf 4452&\it 8638&\bf 1095&\it 2020&\bf 186&\it 329& \bf 30&\it 41&\bf 3&\it 3&\bf 1&\it 1\\ \hline \end{tabular}\vspace*{-0.5cm} \end{table} \section{Conclusions and Outlook} \noindent A prescription for defining a standard basis set of invariants in non--Abelian gauge theories was obtained. A reduction algorithm was presented to convert a given Lorentz scalar by partial integration, by the Bianchi identity, and by cyclic invariance of the trace into a linear combination of this basis set of invariants. The proof that this set is a basis indeed, relies on a graphical representation of invariants and is given elsewhere.\cite{proof} For cases where, in addition, the mirror transformation reduces the number of independent invariants, a general proof of the basis property is still lacking. However at least up to mass dimension 16, it can be shown by counting the invariants that the standard set remains a basis. Another open problem is to take into account additional identities which exist for particular choices of the gauge group representation. \vspace*{-1mm} \nonumsection{Acknowledgements} \noindent The author would like to thank C.~Schubert for indicating the need to clarify this problem and D.~Fliegner for discussions on the reduction algorithm. The proof--reading of the manuscript by D.~Lehner and G.~Weigt is gratefully acknowledged.\vspace*{-1mm} \nonumsection{References} \noindent \vspace*{-4ex}
\section{Introduction} Quarks are not observed as free particles, but only indirectly inside hadrons; this confinement of the quarks is an essential and very intriguing property of QCD, both from a theoretical and from an experimental point of view. Despite a lot of effort, there are still a lot of open questions about the confinement mechanism and confined particles. One of such questions is what the behavior of the full propagator of a confined particle is: e.g. does it have the same kind of analyticity properties as a bare quark propagator? If the full quark propagator has no mass singularity in the timelike region, it can never be on mass-shell and thus never be observed as a free particle \cite{Co80,GoMa89,Gr91,RoWiKr92}. So in this way the absence of a mass singularity implies directly confinement, and thus the analytic structure of the full quark propagator might be connected with confinement. Since confinement is a nonperturbative phenomenon, the analytic properties of the full fermion propagator in a confining theory have to be studied in a nonperturbative way. The Dyson--Schwinger equation is a very powerful tool to study nonperturbative phenomena, and it is commonly used for studying dynamical chiral symmetry breaking, but it can also be useful in studies of confinement \cite{RoWi94}. The usual truncation schemes of the Dyson--Schwinger equation show that the full fermion propagator in QED and QCD has complex branchpoints, instead of the expected mass singularity on the real timelike axis \cite{AtBl79,StCa90,MaHo92,StCa92,Ma94}. Although this phenomenon might be an artifact of the approximations, as believed about 15 years ago when it was first discovered \cite{AtBl79}, it has been suggested more recently that it might be a genuine property of the full theory, connected with confinement, especially in QCD \cite{StCa90,MaHo92,StCa92}. If the quark propagator has a mass-like singularity at complex momenta, instead of a mass singularity in the timelike region, it can never be on mass-shell and is thus confined. Not only in QCD the fermions are confined, also in several other theories there is confinement. Quantum electrodynamics in two space- plus one time-dimension (QED3) is such a theory, with a confining potential for the fermions, at least at the classical level; for the full theory it depends on the behavior of the vacuum polarization \cite{BuPrRo92}. It is also a very interesting model to study dynamical mas generations, and for this purpose the Dyson--Schwinger equation has been extensively studied on the Euclidean axis \cite{Pi84,appetal,DaKoKo89,PeWe88,AtJoMa90,Penetal92}. The theory is super-renormalizable, and does not suffer from the ultraviolet divergences which are present in the corresponding four-dimensional theories. That means that we do not need to introduce any artificial cutoff, and the only mass scale in massless QED3 is the dimensionful coupling. In this way we are provided with a very interesting model, from which we can learn a lot about the analytic structure of the propagator, and which is mathematically easier to analyze than four-dimensional theories. The result can be very useful as guidance for other, more complicated, theories like QCD. Apart from the the interesting features connected with dynamical mass generation and confinement in general, it might also have some direct physical relevance, both in condensed matter physics (in connection with phenomena occurring in planes) and as the high-temperature limit of the corresponding four-dimensional theory. In this paper we study the analytic structure of the fermion propagator in QED3, using the Dyson--Schwinger equation and some different approximations for the full photon propagator. We show that, if there is a confining potential, the fermion propagator has complex mass-like singularities, but if there is no confining potential, the mass singularities are located almost on the real timelike axis, as we would expect. The presence or absence of the confining potential depends on the particular approximation for the photon propagator. This paper is organized as follows: in the next section we review the analytic structure of the fermion propagator in the context of perturbation theory and what we would expect for the propagator of a confined particle. In Sec.~\ref{secform} we introduce the model we are considering and its confining properties. Next, we discuss the Dyson--Schwinger equation, the truncation scheme we are using, and our numerical procedures. In Sec.~\ref{secres} we present our results, and finally we give some conclusions in Sec.~\ref{secconc}. \section{Analytic structure of propagators} \label{secans} \subsection{Mass singularities} The analytic structure of the bare fermion propagator is well-known: in momentum space, it has a single pole at the bare mass of the fermion. In Minkowski metric (which we will use in this section only, out of convenience), we have for the bare propagator \begin{eqnarray} S(p) &=& \frac{1}{\not{\! p} - m_0 + i\varepsilon} \,, \end{eqnarray} with a mass-pole which is located at timelike momentum $p^2_{\hbox{\scriptsize Mink}} = m^2_0$. The integration contour one encounters in all kinds of calculations, goes around this singularity due to the $i\varepsilon$-prescription, and this $i\varepsilon$ also allows us to perform the usual Wick rotation from Minkowski space to Euclidean space. In perturbation theory, the full fermion propagator has a similar structure, at least on the first Riemann sheet: a single pole at the physical mass of the particle, and a more complicated structure for momenta beyond some threshold energy for multi-particle production, see Fig.~\ref{fig1}. If we are dealing with massless particles, as in QED, where we have massless photons, this single pole becomes a logarithmic branchpoint, see e.g. \cite{itzu}. In general, we expect a similar structure for the full fermion propagator in a nonperturbative calculation, at least if the fermion corresponds to a stable physical particle. In a theory of interacting particles {\em with asymptotic states} we have the K\"allen--Lehmann representation \begin{eqnarray} \label{klrep} S_F(p) &=& Z_2 \frac{\not{\! p} + m_{\hbox{\scriptsize phys}}} {p^2 - m_{\hbox{\scriptsize phys}}^2} + \int_{\tilde m^2}^\infty {\rm d}\mu^2 \frac{\not{\! p} \rho_1(\mu^2) + \rho_2(\mu^2)} {p^2 - \mu^2 + i\varepsilon} \,, \end{eqnarray} with the spectral weight functions $\rho_i(p^2)$ real and nonnegative, and $\tilde m \ge m_{\hbox{\scriptsize phys}}$ the threshold for multi-particle production. We therefore expect a full electron propagator with a mass singularity at the physical mass of the electron, which is located on the real axis in the timelike region at $p^2_{\hbox{\scriptsize Mink}} = m^2_{\hbox{\scriptsize phys}}$, and a logarithmic branch-cut along the real axis, beyond this singularity. However, the derivation of the K\"allen--Lehmann representation breaks down in the absence of the asymptotic states; the above argument only holds in cases were the fermion is indeed a stable, physically observable particle. If we are considering a theory with confined fermions, which means that there are no asymptotic states for these fermions, we do not have a rigorous proof of the existence of a K\"allen--Lehmann representation, so we do not know a priori the analytic structure of the propagator of such a confined particle. The mass singularities of the propagator at the physical mass of the particle are crucial for the existence of observable asymptotic states. Without such mass singularities, the particles can never be on mass shell, and thus never be observed as real particles. In other words, confinement might very well be related to the absence of such mass singularities, and thus to the absence of a K\"allen--Lehmann representation for the propagator of such a particle \cite{Co80,GoMa89,Gr91,RoWiKr92}. \subsection{Complex singularities} If the propagator of a confined particle does not have a a K\"allen--Lehmann representation, what (other) analytic structure can we expect? Writing the full fermion propagator as \begin{eqnarray} S(p) &=& Z(p^2)\frac{\not{\! p} + m(p^2)}{p^2 - m^2(p^2)} \,, \end{eqnarray} we can now ask the question: what analytic structure is possible? In principle there are the following possibilities: \begin{itemize} \item the propagator has complex singularities (at zeros of the denominator); \item the propagator is an entire function; \item the propagator has compensating zeros (both the denominator and the wavefunction renormalization $Z(p)$ are zero at the same point, which might be located in the timelike region). \end{itemize} During the last couple of years, analyses of the fermion propagator using the Dyson--Schwinger equation in the complex momentum plane show complex mass-like singularities in a variety of models and truncation scheme. This phenomenon was first discovered by Atkinson and Blatt \cite{AtBl79} in quenched ladder QED4, and it was generally believed to be an artifact of the approximations. However, in a theory with confined particles, it might very well be a genuine property of the full theory: the absence of a mass singularity at timelike momenta will effectively confining the particles, in the sense that they will not be observable as physical stable states. Recently, it has been suggested by several authors, that the complex singularities one finds by solving the Dyson--Schwinger equation for complex momenta, are indeed a signal for confinement, especially in a confining theory like QCD \cite{StCa90,MaHo92,StCa92}. In this paper we show that there is indeed a connection between a confining potential and complex mass-like singularities in QED3. Note however that also other analytic structures, like a fermion propagator which is an entire function, will effectively confine the fermions; and in principle there are also other confinement mechanisms possible which do allow for a physical mass pole for the fermion propagator. \section{QED3} \label{secform} \subsection{Formalism} In Minkowski space, we need the artificial $i\varepsilon$ description, in order to define the path integrals, and to select the integration path around the mass singularities in all kinds of calculations. Alternatively, we could set up our field theory in Euclidean space, in which case the integrals are well-defined from the beginning. In principle the Wick rotation allows us to go from Euclidean to Minkowski space and back, and both formulations seem to be equivalent, but in the presence of complex singularities this easy connection between Euclidean and Minkowski space is destroyed. Since the theory is better defined in Euclidean metric, we will use that formalism. Once we know the Euclidean Green's functions, we can obtain the Wightman functions in coordinate space by an analytic continuation in the time-coordinates, and from them the physically relevant Minkowski Green's functions \cite{glijaf}. In this way we can (in principle) extract all of the physically relevant information in Minkowski space, even after setting up the formalism in Euclidean space. The Lagrangian in Euclidean space is \begin{eqnarray} {\cal L}(\psi, \bar\psi, A) &=& \bar\psi \left( \gamma^\mu(\partial^\mu + ieA^\mu) + m_0 \right) \psi + \textstyle{\frac{1}{4}} F^{\mu\nu}F^{\mu\nu} + \textstyle{\frac{1}{2a}} (\partial^\mu A^\mu)^2 \,. \end{eqnarray} In QED3, only three anti-commuting $\gamma$-matrices are needed, which can be realized by taking a two-dimensional representation for these matrices, e.g. the Pauli spin matrices. In that case we also have two-dimensional spinors, instead of the four-dimensional spinors one would use in four-dimensional theories. However, here we will use the formulation with a four-dimensional spinor-space, and use the same $\gamma$-matrices as in four space-time dimensions. In order to study dynamical mass generations, and its influence on confinement, we take the bare fermions to be massless: $m_0 = 0$. In QED3 with four-dimensional spinors one can have two types of mass terms for the fermions, namely a parity-breaking and a parity-conserving mass term. We will only consider the dynamical generation of a parity-even mass \cite{appetal}; it has been shown that there is no dynamical breakdown of parity \cite{VaWi84}. The full fermion propagator can be written as \begin{eqnarray} S^{-1}(p) &=& Z(p) \left( i\not{\! p} + m(p) \right) \,, \end{eqnarray} and the full photon propagator is \begin{eqnarray} \label{photprop} D^{\mu\nu}(q) &=& \frac{1}{q^2(1 + \Pi(q))} \left( \delta^{\mu\nu} - \frac{q^\mu q^\nu}{q^2} \right) + a \frac{q^\mu q^\nu}{q^4} \,, \end{eqnarray} in a general covariant gauge. In this equation, $\Pi(q)$ is the vacuum polarization, $a$ the gauge parameter, $m(p)$ the dynamical mass function of the fermion, and $Z(p)$ the fermion wavefunction renormalization. For sake of simplicity, we use the Landau gauge ($a=0$). The exact Dyson--Schwinger equation for the fermion propagator is \begin{eqnarray} \label{dseqn} S^{-1}(p) &=& i\not{\! p} + e^2 \int\frac{{\rm d}^3k}{(2\pi)^3} \gamma^\mu S(p) \Gamma^\nu(p,k) D^{\mu\nu}(p-k) \,, \end{eqnarray} with the unknown full vertex $\Gamma^\nu(p,k)$, and the full photon propagator $D_{\mu\nu}(p-k)$. In analyzing the fermion Dyson--Schwinger equation, we have to truncate this equation. In this paper, we discuss both the socalled quenched ladder approximation (bare photon and bare vertex), and two approximations based on the $1/N$ expansion. \subsection{$1/N$ Expansion} A very popular truncation scheme in QED3 is the $1/N$ expansion: consider $N$ massless fermion flavors, and use the large $N$ limit in the following way: let $N \rightarrow \infty$ and $e^2 \rightarrow 0$ in such a way that the product $e^2 N$ remains fixed. It has been shown that massless QED3 is infrared finite order by order in such a $1/N$ expansion \cite{expansion}. For convenience we choose the coupling, which defines our mass scale, to be \begin{eqnarray} e^2 &=& \frac{8 \alpha}{N} \,, \end{eqnarray} and keep $\alpha$ fixed. Such a $1/N$ expansion means that we have to take into account the one-loop vacuum polarization: the coupling is of order $1/N$, but there are $N$ fermion loops contributing to the vacuum polarization tensor \begin{eqnarray} \label{vacpol} \Pi^{\mu\nu}(q) &=& - e^2 \, N \int\frac{{\rm d}^3k}{(2\pi)^3} {\rm Tr}\left[ \gamma^\mu S(k+q) \Gamma^\nu(k+q,k) S(k) \right] \,. \end{eqnarray} This vacuum polarization tensor has an ultraviolet divergence in its part proportional to $\delta^{\mu\nu}$, which can be removed by evaluating it using a gauge-invariant regularization scheme. Defining the vacuum polarization $\Pi(q)$ by \begin{eqnarray} \Pi^{\mu\nu}(q) &=& \left(q^2 \delta^{\mu\nu} - q^\mu q^\nu \right) \Pi(q) \,, \end{eqnarray} we can get the regularized vacuum polarization by contracting the vacuum polarization tensor with \cite{RoWi94,BuPrRo92} \begin{eqnarray} \left( q^2 \delta^{\mu\nu} - 3 q^\mu q^\nu \right) /q^4 \,, \end{eqnarray} which is orthogonal to $\delta^{\mu\nu}$ and thus projects out the divergent part. Note that this regularization gives the same result as dimensional regularization, but this projection is much easier to perform if we take into account dynamical fermions, in Sec.~\ref{subsecdyn}. Using bare, massless fermions and a bare vertex, we have for this vacuum polarization \begin{eqnarray} \label{vacpolml} \Pi(q) &=& \frac{e^2 \, N}{8 \sqrt{q^2}} \;=\; \frac{\alpha}{q} \,, \end{eqnarray} whereas the one-loop vacuum polarization with massive fermions gives \begin{eqnarray} \label{vacpolmass} \Pi(q) &=& \frac{2\alpha}{\pi \, q^2} \left( 2m + \frac{q^2 - 4m^2}{q} \arcsin{\frac{q}{\sqrt{q^2 + 4m^2}}} \right) \,. \end{eqnarray} The crucial difference between the vacuum polarization with massless and with massive fermions lies in the infrared region: with massless fermions the vacuum polarization blows up at the origin, $\Pi(0) \rightarrow \infty$. With massive fermions however, with a constant mass $m$, the vacuum polarization is finite in the infrared \begin{eqnarray} \label{vacpolmassir} \Pi(0) &\rightarrow& \frac{4\alpha}{3\pi \, m} \,. \end{eqnarray} \subsection{Confining potential} One of the interesting properties of QED3 is that it exhibits confinement \cite{GoMa82}. We can define a ``classical'' potential for the fermions in coordinate space \cite{BuPrRo92} \begin{eqnarray} V(\vec{x}) &=& -e^2 \int\frac{{\rm d}^2q}{(2\pi)^2} e^{i\vec{q}\cdot\vec{x}} \frac{1}{\vec{q}^2(1+\Pi(\vec{q}^2))} \,, \end{eqnarray} where $\Pi(q)$ is the vacuum polarization. In lowest order in perturbation theory, we can neglect the effects of the vacuum polarization and simply calculate the potential. This leads to a logarithmically rising potential \begin{eqnarray} V(\vec{x}) &=& \frac{e^2}{2\pi} \ln(e^2 |x|) \,. \end{eqnarray} Because this potential increases at large distances, it effectively confines the fermions, and there are no free asymptotic one-fermion states possible. Of course, this potential will change under the influence of the vacuum polarization. As our results show, the correct inclusion of the vacuum polarization is indeed essential for confinement. The relevant region for the question whether or not there is a confining potential is the infrared region, corresponding to large spatial separations in configuration space. As is shown by Burden {\em et al.} \cite{BuPrRo92}, under quite general and natural conditions for the vacuum polarization, the potential associated with the full photon propagator behaves like \begin{eqnarray} V(\vec{x}) &=& \frac{e^2\,\ln{\big(e^2 |x|\big)} }{(1 + \Pi(0))\,2\pi} + \hbox{constant} + {\cal O}\left(\frac{1}{|x|}\right) \,. \end{eqnarray} This behavior can be derived assuming that the vacuum polarization is bounded and continuously differentiable for Euclidean momenta, and that it falls of at least as $1/q$ as $q \rightarrow \infty$. {}From this equation, we can see immediately that, depending on the behavior of the vacuum polarization in the infrared region, there are two possibilities \begin{itemize} \item a confining potential if $\Pi(0)$ is finite; \item no confining potential if $\Pi(0) \rightarrow \infty$. \end{itemize} If we know look at the one-loop perturbative vacuum polarization, we see that there is an essential difference between massless and massive QED3. For massive fermions, the vacuum polarization at the origin, $\Pi(0)$, is finite as can be seen from Eq.~(\ref{vacpolmassir}). Therefore there is a logarithmically confining potential in leading order in $1/N$ for massive fermions. On the other hand, if the fermions are massless, there is no confining potential to leading order to $1/N$; the one-loop vacuum polarization blows up at $q^2 \downarrow 0$, see Eq.~(\ref{vacpolml}), and the photon propagator is softened in the infrared region. Perturbatively, to leading order in $1/N$, the fermion propagator is just the bare propagator, with a single pole at the origin, corresponding to an observable massless fermion. However, a dynamically generated fermion mass might very well change this leading-order behavior. \section{Dyson--Schwinger equation} \label{secdse} In order to determine the analytic structure of the fermion propagator nonperturbatively, we use the Dyson--Schwinger equation. In general, after reducing the $\gamma$-algebra, the Dyson--Schwinger equation, Eq.~(\ref{dseqn}), becomes \begin{eqnarray} \label{dsmassgen} Z^{-1}(p)\,m(p) &=& e^2 \int\frac{{\rm d}^3k}{(2\pi)^3} \textstyle{\frac{1}{4}} {\rm Tr}[\gamma^\mu S(k) \Gamma^\nu(p,k) D^{\mu\nu}(q)] \,, \\ % \label{dswavegen} Z^{-1}(p) & = & 1 - \frac{e^2}{p^2} \int\frac{{\rm d}^3k}{(2\pi)^3} \textstyle{\frac{1}{4}} {\rm Tr}[\not\! p \gamma^\mu S(k) \Gamma^\nu(p,k) D^{\mu\nu}(q)] \,, \end{eqnarray} with the photon propagator $D^{\mu\nu}(q)$ as defined by Eq.~(\ref{photprop}), with the unknown vacuum polarization, and the unknown full vertex function $\Gamma^\nu(p,k)$. The general approach to solve this equation is to choose a specific truncation scheme for the vertex and the photon propagator, and then to solve the resulting equations numerically. \subsection{Truncation scheme} Here, we will adopt the bare vertex approximation, replacing the full vertex by the bare one, $\gamma^\mu$. We also neglect the effects of the wavefunction renormalization, so we put $Z(p) = 1$. This truncation is based on the leading-order behavior of the vertex and the wavefunction renormalization in the $1/N$ expansion. Such an approximation scheme is also consistent with the requirement following from the Ward--Takahashi identity that the wavefunction renormalization and the vertex renormalization are equal. It is usually referred to as the ladder or rainbow approximation, and it leads to a finite critical number of fermion flavors below which the chiral symmetry is broken dynamically \footnote{Using this approximation, the equation for the wavefunction renormalization, Eq.~(\ref{dswavegen}), is formally satisfied up to order $1/N$. It is known that the effects of the wavefunction renormalization (together with a more sophisticated Ansatz for the vertex) will change the results found in this $1/N$ truncation scheme \cite{PeWe88,AtJoMa90,Penetal92}, but we will not address that problem here.}. Note that in the Landau gauge, with a bare photon propagator and bare vertex (the quenched ladder approximation), Eq.~(\ref{dswavegen}) gives $Z(p)=1$ exactly. For the photon propagator we use some different approximations, to determine the influence of the infrared behavior of the vacuum polarization on the analytic structure of the fermion propagator and on the (confining) potential. We compare in detail the results as obtained in \begin{enumerate} \item the quenched approximation: a bare photon propagator; \label{trunc1} \item the $1/N$ expansion using the analytical formula for the one-loop vacuum polarization with bare, massless fermions, Eq.~(\ref{vacpolml}) \label{trunc2} \item the $1/N$ expansion using the one-loop vacuum polarization with full fermions, {\em with the dynamically generated fermion mass function}. \label{trunc3} \end{enumerate} In both \ref{trunc2} and \ref{trunc3}, we take a bare vertex in the expression for the vacuum polarization. This truncation scheme gives us the following expression for the mass function \begin{eqnarray} \label{genmasseq} m(p) &=& \frac{e^2}{2\,\pi^2} \int_0^\infty {\rm d}k \frac{k^2 \, m(k)}{k^2 + m^2(k)} {\rm K}(p,k) \,, \\ % {\rm K}(p,k) &=& \int_0^\pi \frac{\sin{\theta} \, {\rm d}\theta} {(p^2-2pk\cos{\theta}+k^2)(1+\Pi(p^2-2pk\cos{\theta}+k^2))} \,, \end{eqnarray} with the kernel ${\rm K}(p,k)$ depending on the particular approximation we use for the photon propagator. \subsection{Numerical calculations} Once we have truncated the equations, we can solve the resulting integral equation for the mass function numerically. We start by solving the equation for Euclidean momenta $0 \leq p^2 < \infty$. However, we are not really interested in the result on the Euclidean axis (for a more detailed discussion about the existence of a critical number of fermion flavors for dynamical chiral symmetry breaking we refer to the literature \cite{appetal,DaKoKo89,PeWe88,AtJoMa90,Penetal92}), but we want to know the behavior of the propagator in the complex momentum plane. For that purpose we have used two different approaches: one is a direct analytic continuation of the integral equation, Eq.~(\ref{genmasseq}), into the complex plane. This can be done by deforming the integration contour and solving the integral equation along this new contour. Note that it is not possible to keep the integration variable $k$ real, and take only the external variable $p$ complex (after solving the integral equation on the real axis), because of the analytic structure of the kernel ${\rm K}(p,k)$. With massless photons, and thus a photon propagator which has a singularity at the origin, there is a pinch singularity at $p=k$, and we are forced to integrate through the point $p=k$. So for complex momenta $p$ we have to solve the integral equation along a deformed contour in the complex plane. In practice, we change the integration contour by rotating it in the complex plane, multiplying both the internal and the external variable by a phase factor ${\rm e}^{i\phi}$, so we get the complex variables $\tilde k = {\rm e}^{i\phi} k$ and $\tilde p = {\rm e}^{i\phi} p$, see Fig.~\ref{fig2}. Since in QED3 the integral falls off rapidly enough in the ultraviolet, there is no need to take into account the contribution coming from the arc at infinity, in contrast to theories with a finite cutoff like QED4. This procedure works quite well, until one comes close to a singularity caused by a zero of the denominator of the integration kernel \begin{eqnarray} \frac{1}{k^2 + m^2(k)} \,, \end{eqnarray} where the numerical integration procedure becomes unstable. The location in the complex plane of the actual singularity itself can be obtained by extrapolating the numerical results to the ``physical'' mass $\mu$, defined by the zero of this denominator \begin{eqnarray} - \mu^2 + m^2(\sqrt{-\mu^2}) &=& 0 \,. \end{eqnarray} For more details about our numerical procedure and the analytic continuation, we refer to \cite{pmthesis}. In this way we can in principle find the singularities in the complex plane, but it is a very time-consuming numerical process, and does not always converge to a stable solution. Therefore we also used another method, based on the Euclidean-time Schwinger function, to determine whether or not the propagator corresponds to a physical observable state \cite{RoWi94,HoRoMc92}. We define \begin{eqnarray} \sigma(p^2) &=& \frac{m(p)}{p^2 + m^2(p)} \,, \\ \Delta(t) &=& \int{\rm d}^2 x \int \frac{{\rm d}^3p}{(2\pi)^3} {\rm e}^{i(p_3 t + \vec{p}\cdot\vec{x})} \sigma(p^2) \,. \end{eqnarray} Using this Schwinger function, one can show that if there is a stable asymptotic state associated with this propagator, with a mass $m$, then \begin{eqnarray} \Delta(t) &\sim& {\rm e}^{- m t} \end{eqnarray} for large (Euclidean) $t$, so for the logarithmic derivative we get \begin{eqnarray} \lim_{t \rightarrow\infty} \frac{\rm d}{{\rm d}t} \ln\left(\Delta(t)\right) &=& - m \,, \end{eqnarray} whereas two complex conjugate mass-like singularities, with complex masses $\mu = a \pm i\,b$, lead to an oscillating behavior like \begin{eqnarray} \Delta(t) &\sim& {\rm e}^{- a t} \cos{(bt + \delta)} \end{eqnarray} for large $t$. This method is much less time-consuming to see whether or not the propagator has a real mass-singularity or not, but it is less accurate than solving the Dyson--Schwinger equation for complex momenta in determining the (complex) mass-singularities. \section{Results} \label{secres} \subsection{Quenched QED3} \label{subsecquen} In massless quenched QED3, there is no free parameter: the coupling in QED3 is dimensionful and thus defines the energy scale, and there are no other parameters. By choosing the Landau gauge, we satisfy the requirement that the wavefunction renormalization and the vertex renormalization are exactly equal: from Eq.~(\ref{dswavegen}) it follows directly that in quenched QED using the Landau gauge, $Z(p)=1$, and the equation for the mass function reduces to \begin{eqnarray} m(p) &=& \frac{e^2}{2\pi^2} \int_0^\infty {\rm d}k \frac{m(k)}{k^2 + m^2(k)} \frac{k}{2p} \ln{\frac{(p+k)^2}{(p-k)^2}} \,. \end{eqnarray} Solving this equation on the Euclidean axis shows that there is dynamical mass generation in this case, and the infrared mass $m(0)$ is proportional to the dimensionful coupling, as expected. Next, we have calculated the Schwinger function, using the mass function on the Euclidean axis, see Fig.~\ref{fig3}. This figure clearly shows that there is no stable asymptotic one-fermion state associated with this propagator, in other words, the fermions cannot be observed as free particles and are thus confined. The oscillations in this figure strongly suggest that the fermion propagator has complex mass-like singularities, corresponding to two complex-conjugate masses. Using \begin{eqnarray} \label{schwcmplxmass} \Delta(t) &\sim& {\rm e}^{- a t} \cos{(bt + \delta)} \,, \end{eqnarray} to extract such a complex mass, we estimate this to be \begin{eqnarray} \mu &=& (0.80 \pm 0.71 \, i)\,m(0) \end{eqnarray} However, this method might be not very accurate in determining the actual value of the (complex) masses, since Eq.~(\ref{schwcmplxmass}) only holds for large values of $t$, whereas for large values of $t$ the numerical noise in calculating the Euclidean-time Schwinger function destroys the signal. So we have used values of $t$ up to $t \sim 10/m(0)$ (we rescale all dimensionful quantities by $m(0)$), and use only the first oscillations to determine the imaginary part of the complex mass. Furthermore, Eq.~(\ref{schwcmplxmass}) is based on the assumption that only these (complex) singularities contribute to the Schwinger function (at least at large $t$), but if there are complex singularities, there might be more than just two complex-conjugate mass-like singularities. Therefore we also used the other method to determine the analytic structure of the propagator, and have solved the integral equation in the complex plane. This leads to two complex-conjugate singularities, located at \begin{eqnarray} |\mu| &=& 0.104 \,e^2 \;=\; 1.01\,m(0) \,, \\ \theta_\mu &=& {\textstyle{\frac{\pi}{2}}} - \phi \; = \; 0.819 \,. \end{eqnarray} This result confirms the estimate based on the Schwinger function, given the inaccuracy of the estimate of the complex mass. So both the Schwinger function and a direct search for mass-like singularities show that there is a complex mass singularity, which makes it impossible for the fermion propagator to become on mass-shell, end thus effectively confines the fermion. This is in agreement with the fact that in quenched massless QED3 there is a confining potential. \subsection{One-loop vacuum polarization} \label{subsecbare} Next, we include the one-loop vacuum polarization, using bare massless fermions, Eq.~(\ref{vacpolml}). As already mentioned before, perturbatively the $1/N$ expansion gives to leading order no confining potential, and a full fermion propagator which is the same as the bare one, and thus corresponding to a massless stable asymptotic state. In the case of dynamical mass generation, which we consider here, the situation is more complicated. Since the number of fermion flavors is the only free parameter in this case, we present our results as a function of $N$. For simplicity we use Landau gauge, as in the quenched approximation, and we can perform the angular integration in the Dyson--Schwinger equation analytically to arrive at the equation for the mass function \begin{eqnarray} \label{masseqnml} m(p) &=& \frac{4\,\alpha}{N\,\pi^2} \int_0^\infty {\rm d}k \frac{m(k)}{k^2 + m^2(k)} \frac{k}{p} \ln{\frac{|p+k|+\alpha}{|p-k|+\alpha}} \,, \end{eqnarray} This can be solved numerically as an integral equation, or after expanding the logarithm and some further approximations reduced to a second-order nonlinear differential equation \cite{appetal}. Both the integral and the differential equation show that there is dynamical mass generation if $N < N_c = 3.24$, see Fig.~\ref{fig4mass}; this critical number is in agreement with analytical calculations using bifurcation theory, leading to $N_c = 32/\pi^2$. We have calculated the Schwinger function, using the mass function on the Euclidean axis, see Fig.~\ref{fig3}. This figure strongly suggest that there is a stable asymptotic one-fermion state, which means that the fermions are not confined and can be observed. Up to the largest values of $t$ at which the Schwinger function gives a numerically stable result, we find an almost constant logarithmic derivative. {}From this Schwinger function we have derived a value for the asymptotic mass for some different number of fermion flavors, and the results are listed in Table~\ref{table}. We do not find any evidence for oscillations which would signal a complex mass, as there were in the quenched approximation. We have also solved the integral equation in the complex plane. This reveals that the mass singularities are not exactly on the real timelike axis, but that they do have small imaginary parts, see Table~\ref{table}. In Fig.~\ref{fig5phase} we have plotted the phase of these singularities as a function of $N$, by numerically solving both the integral equation and the differential equation which can be derived from Eq.~(\ref{masseqnml}), and those results almost coincide. Given the fact that we have only solved the truncated Dyson--Schwinger equation, it is not unreasonable to expect a small deviation of the physical mass from the real axis, and given the relative smallness of the imaginary part this could very well be an artifact of the approximations, especially since the singularity tends to move toward the real timelike axis if the number of flavors goes to the critical number. The reason for not finding this imaginary part of the mass singularities using the Schwinger function lies in its smallness: since the imaginary part is of the order of $10\%$ of the real part (or even less), we will not find a clear signal for it at $t \leq 10/m$, where $m$ is the typical infrared mass-scale; however, the numerical noise destroys the signal completely at these (or larger) values of $t$. So our conclusion is that this approximation, using the one-loop vacuum polarization of bare massless fermions, leads to (almost) stable observable asymptotic states, with an (almost) real physical mass. This agrees well with the fact that in this case we do not have a confining potential. \subsection{Full vacuum polarization} \label{subsecdyn} Finally, we include the one-loop vacuum polarization, Eq.~(\ref{vacpol}), with dynamical fermions and a bare vertex. In other words, we consider the coupled Dyson--Schwinger equations for the photon and propagator, in the bare vertex approximation. This leads to two coupled integral equations to solve \begin{eqnarray} m(p) &=& \frac{4\,\alpha}{N\,\pi^2} \int_0^\infty {\rm d}k \frac{k^2 \, m(k)}{k^2 + m^2(k)} \int_{-1}^{1}\frac{{\rm d}z}{(p^2-2pkz+k^2)(1+\Pi(p^2-2pkz+k^2))} \,, \\ \Pi(q^2) &=& \frac{8\alpha}{q^2} \int\frac{{\rm d}^3k}{(2\pi)^3} \frac{2 k^2 - 4 k\cdot q - 6 (k\cdot q)/q^2} {(k^2 + m^2(k))((k+q)^2 + m^2(k+q))} \,. \end{eqnarray} Again, this can be solved numerically: we start by solving the mass equation for a given vacuum polarization, and use that resulting mass function to calculate the vacuum polarization numerically and iterate this procedure. Just as in the previous case, it leads to dynamical mass generation if the number of fermion flavors is below a critical number, see Fig.~\ref{fig4mass}. The behavior of the infrared mass is quite similar, and also the critical number is the same as in the previous approximation, $N_c = 3.24$, as could be expected on grounds of bifurcation theory. Also in this case we have calculated the Schwinger function, see Fig.~\ref{fig3extra}. This shows that there are no stable asymptotic one-fermion states associated with this propagator, just as in quenched QED3, but in sharp contrast to the previous case. We have also shown the result with a fixed mass ($m=m(0)$) in the analytical formula for the vacuum polarization, Eq.~(\ref{vacpolmass}). This gives qualitatively the same result as when using the dynamical mass function. In both cases the oscillations indicate complex mass-like singularities, and we have given estimates for these complex masses in Table~\ref{table}. We have also solved the integral equation in the complex plane, which confirms the observation based on the Schwinger function that there are complex mass-like singularities. The phase of these singularities is plotted in Fig.~\ref{fig5phase}, and for some different values of $N$ we have given our result in Table~\ref{table}. Given the inaccuracy in the estimates based on the Schwinger function, there is a good agreement between both methods. It is clear that the the effects of the fermion mass in the loop for the vacuum polarization confines the fermions: the potential becomes confining, the mass singularities move into the complex plane, and there are no stable observable asymptotic states. \subsection{Discussion of the results} \label{subsecdisc} Our results show that the correct treatment of the vacuum polarization is essential in a nonperturbative calculation of the fermion propagator. Based on bifurcation theory, one can argue that the influence of the dynamically generated mass function can be neglected in studying the chiral phase transition. Although this is indeed true for the value of the critical coupling, and maybe also for the behavior of the infrared mass $m(0)$ close to the critical coupling, it is certainly not true for the behavior of the {\em physical} mass, defined at the zero of $p^2 + m^2(p)$. On the real axis, the dynamical mass function in quenched QED3 is qualitatively quite similar to the mass function in the $1/N$ expansion, both with massless fermions and with massive fermions in the vacuum polarization, see Fig.~\ref{fig6}(a). There is a scale difference between the different approximations, but all mass functions are almost constant in the (far) infrared region, and fall of to zero as $1/p^2$ in the (far) ultraviolet. Only in the intermediate-energy region there are some differences due to the inclusion of the vacuum polarization. In contrast, in the complex plane the behavior is not similar at all, leading to a drastic different analytic structure. This difference can be traced back to the difference in the infrared behavior of the photon propagator: with a confining photon propagator there are complex mass-like singularities, whereas with a deconfining photon propagator these singularities are located almost on the real timelike axis. Surprisingly, this difference can be seen very clearly by using the Euclidean-time Schwinger function, which can be calculated using the mass function on the real Euclidean axis only. So although the behavior of the mass function in the Euclidean region looks quite similar, there are essential differences which can be shown explicitly by calculating this Schwinger function. This means that this method is indeed a useful way to determine whether the propagator corresponds to a confined particle or to a physical observable particle. The difference in analytic structure is due to the difference in the infrared behavior of the photon, or more precisely, due to the different behavior of $\Pi(0)$ in the different approximations. In Fig.~\ref{fig6}(b) we have plotted both the one-loop vacuum polarization, for massless and massive fermions, and the full vacuum polarization calculated numerically with dynamical massive fermions. Only in the infrared region there is a difference, and it is exactly this difference that causes the different behavior of the fermion propagator in the complex plane; it is also this infrared behavior which makes the potential confining or not. Therefore our conclusion is that (at least in this model) confinement is caused by the infrared behavior of the photon propagator, and is connected with complex mass-like singularities of the fermion propagator, thus preventing the fermions from being on mass-shell. Finally, we should remark that these calculations are all done in the bare vertex approximation in the Landau gauge. It is known that the effects of vertex corrections, together with the wavefunction renormalization $Z(p)$ which we have set equal to $1$, can change the results quite drastically \cite{PeWe88,AtJoMa90,Penetal92}. Another question is what happens in other gauges, whether or not our results are gauge independent. As a qualitative indication whether or not our conclusions about confinement in QED3 also hold beyond the bare vertex approximation and in other gauges, we could compare our numerical vacuum polarization with the vacuum polarization as obtained by Burden {\em et al}. They solved the Dyson--Schwinger equation for the fermion propagator in quenched QED3 with the Ball--Chiu Ansatz for the vertex, and used this propagator to calculate the vacuum polarization, again with the Ball-Chiu vertex. Qualitatively our result for the vacuum polarization agrees with theirs, both in the infrared region, where we find a finite value of $\Pi(0)$ if we take into account the fermion mass, and in the ultraviolet region. In the infrared region there is a quantitative difference, but this can be explained by the fact that the value of $\Pi(0)$ strongly depends on the infrared value of the mass function $m(0)$, which is quite different in different approximations. Of course, we should keep in mind that if the behavior looks similar on the real axis, it does not necessarily mean that they are indeed similar in the entire complex plane. However, the fact that they also found a finite value of $\Pi(0)$ indicates that also beyond the bare vertex approximation there is a confining potential, and we would expect complex singularities as well. Whether or not these singularities are gauge independent (with a suitable vertex Ansatz), will be addressed in the future. Note that also the vacuum polarization itself should be explicitly gauge-independent. \section{Conclusions} \label{secconc} Our results show very clearly that there is a relation between a confining potential, the absence of stable asymptotic states, and complex mass-like singularities. Both in quenched QED3, and in massive QED3 using the $1/N$ expansion (with a dynamically generated fermion mass), there is a logarithmically confining potential, and we show that there are no stable asymptotic states. The Euclidean-time Schwinger function has an oscillatory behavior in these cases, indicating complex mass-like singularities. By solving the Dyson--Schwinger equation directly in the complex momentum plane, we show that there are indeed such complex singularities. On the other hand, using massless bare fermions in the one-loop vacuum polarization, there is no confining potential. In this approximation, the Schwinger function indicates a stable asymptotic state. A direct analysis of the Dyson--Schwinger equation in the complex plane reveals that there are complex singularities even in this case, but that they are located very close to the real timelike axis. Given the approximations made, it is not unreasonable to assume that this small (less than $10\%$) deviation from the real timelike axis is caused by the truncation of the Dyson--Schwinger equation. These results are obtained with a dynamically generated fermion mass, starting with massless bare fermions. For $N > N_c$, in the chirally symmetric phase, there is no dynamical fermion mass, and the fermion propagator has a singularity at the origin, just as the bare one. This also agrees with the fact that in there is no confining potential in this massless phase (in the $1/N$ expansion), due to the infrared softening of the photon propagator in the presence of massless fermions. Thus the chiral phase transition is a confining phase transition as well, at least in this model. Interpreting the absence of a mass singularity on the real axis in the timelike region as confinement does not completely explain the phenomenon of complex mass-like singularities. If they are indeed a genuine property of the full theory, it leads automatically to confinement, but it has more consequences. One of the consequences is that the naive Wick rotation is not allowed, and one should take into account the contributions coming from the complex singularities in going from Euclidean to Minkowski metric (and back). Another problem is connected with questions of unitarity an causality; however, one should keep in mind that these are requirements for the S-matrix of physical processes, and not necessarily for the propagator of an unphysical (confined) particle. Another question is whether or not these complex singularities have a physical interpretation. Naively, the real part (or the absolute value) could be interpreted as the ``constituent'' mass, and the imaginary part as some ``hadronization length'', in terms of QCD and quark confinement. Such an interpretation is analogous to the interpretation of the poles of instable particles in terms of mass and decay width. A crucial requirement for such an interpretation is that the singularities are gauge independent, which has to be studied in detail. \section*{Acknowledgments} I would like to thank Qing Wang, Craig Roberts, and David Atkinson for useful comments. This work has been financially supported by the Japanese Society for the Promotion of Science.
\section{Introduction} The Standard Model involves three gauge couplings $g_i$, ($i=1,2,3$), a Higgs mass parameter $\mu$ and quartic coupling $\lambda$, and the fermion Yukawa couplings. In the Standard Model Lagrangian one must specify the three Yukawa matrices $\lambda^E_{ij}$, $\lambda^D_{ij}$, $\lambda^U_{ij}$, corresponding to up to 54 real parameters, which after diagonalisation lead to 9 physical masses (6 quark masses and 3 charged lepton masses) and 4 physical quark mixing parameters. Thus the fermion sector of the Standard Model involves either 54 or 13 unknown free parameters, depending on how you choose to count them. Either way, from a fundamental point of view the situation is unacceptable and the fermion mass problem, as it has been called, is one motivation for going beyond the Standard Model. The big question of course is what lies beyond it? We have not yet experimentally studied the mechanism of electroweak symmetry breaking, so one might argue that it is premature to study the fermion mass problem. Unless we can answer this, we have no hope of understanding anything about fermion masses since we do not have a starting point from which to analyse the problem. However LEP has taught us that whatever breaks electroweak symmetry must do so in a way which very closely resembles the standard model. This observation by itself is enough to disfavour many dynamical models involving large numbers of new fermions. By contrast the minimal supersymmetric standard model (MSSM) mimics the standard model very closely. Furthermore, by accurately measuring the strong coupling constant, LEP has shown that the gauge couplings of the MSSM merge very accurately at a scale just above $10^{16}$ GeV, thus providing a hint for possible unification at this scale. On the theoretical side, supersymmetry (SUSY) and grand unified theories (GUTs) fit together very nicely in several ways, providing a solution to the technical hierarchy problem for example. When SUSY GUTs are extended to supergravity (SUGRA) the beautiful picture of universal soft SUSY breaking parameters and radiative electroweak symmetry breaking via a large top quark yukawa coupling emerges. Finally, there is an on-going effort to embed all of this structure in superstring models, thereby allowing a complete unification, including gravity. Given the promising scenario mentioned above, it is hardly surprising that many authors have turned to the SUSY GUT framework as a springboard from which to attack the problem of fermion masses \cite{Rabyi}. Indeed in recent years there has been a flood of papers on fermion masses in SUSY GUTs. Although the approaches differ in detail, there are some common successful themes which have been known for some time. For example the idea of bottom-tau Yukawa unification in SUSY GUTs \cite{btauold} works well with current data \cite{btaunew}. A more ambitious extension of this idea is the Georgi-Jarlskog (GJ) ansatz (see later) which provides a successful description of all down-type quark and charged lepton masses \cite{GJ}, and which also works well with current data \cite{DHR}. In order to understand the origin of the texture zeroes, one must consider the details of the model at the scale $M_X \sim 10^{16}$ GeV. The alternative is to simply make a list of assumptions about the nature of the Yukawa matrices at $M_X$ \cite{frogetal}. For example Ramond, Roberts and Ross (RRR) \cite{RRR} assumed symmetric Yukawa matrices at $M_X$, together with the GJ ansatz for the lepton sector. It is difficult to proceed beyond this without specifying a particular model. Indeed, this model dependence may be a good thing since it may mean that the fermion mass spectrum at low energies is sensitive to the theory at $M_X$, so it can be used as a window into the high-energy theory. Therefore in what follows we shall restrict ourselves to a very specific gauge group at $M_X$. Twenty-one years ago Pati and Salam proposed a model in which the standard model was embedded in the gauge group SU(4)$\otimes$SU(2)$_L \otimes $SU(2)$_R$ \cite{pati}. More recently a superpersymmetric (SUSY) version of this model was proposed in which the gauge group is broken at $M_{X}\sim 10^{16}$ GeV \cite{leo1}. The model \cite{leo1} does not involve adjoint representations and later some attempt was made to derive it from four-dimensional strings, although there are some difficulties with the current formulation \cite{leo2}. The absence of adjoint representations is not an essential prerequisite for the model to descend from the superstring, but it leads to some technical simplifications. Also in the present model, the colour triplets which are in separate representations {}from the Higgs doublets, become heavy in a very simple way so the Higgs doublet-triplet splitting problem does not arise. These two features (absence of adjoint representations and absence of the doublet-triplet splitting problem,) are shared by flipped SU(5)$\otimes $U(1) \cite{flipped}, which also has a superstring formulation. Although the present model and flipped SU(5)$\otimes $U(1) are similar in many ways, there are some important differences. Whereas the Yukawa matrices of flipped SU(5)$\otimes$U(1) are completely unrelated at the level of the effective field theory at $M_{X}$ (although they may have relations coming from the string model) in the present model there is a constraint that the top, bottom and tau Yukawa couplings must all unify at that scale. In addition there will be Clebsch relations between the other elements of the Yukawa matrices, assuming they are described by non-renormalisable operators, which would not be present in flipped SU(5)$\otimes $U(1). In these respects the model resembles the SO(10) model recently analysed by Anderson et al \cite{Larry}. However it differs from the SO(10) model in that the present model does not have an SU(5) subgroup which is central to the analysis of the SO(10) model. In addition the operator structure of the present model is totally different. Thus the model under consideration is in some sense similar to flipped SU(5)$\otimes$U(1), but has third family Yukawa unification and precise Clebsch relationships as in SO(10). We find this combination of features quite remarkable, and it seems to us that this provides a rather strong motivation to study the problem of fermion masses in this model \cite{422}. \section{The 422 Model} Here we briefly summarise the parts of the model which are relevant for our analysis. The gauge group is, \begin{equation} \mbox{SU(4)}\otimes \mbox{SU(2)}_L \otimes \mbox{SU(2)}_R. \label{422} \end{equation} The left-handed quarks and leptons are accommodated in the following representations, \begin{equation} {F^i}^{\alpha a}=(4,2,1)= \left(\begin{array}{cccc} u^R & u^B & u^G & \nu \\ d^R & d^B & d^G & e^- \end{array} \right)^i \end{equation} \begin{equation} {\bar{F}}_{x \alpha}^i=(\bar{4},1,\bar{2})= \left(\begin{array}{cccc} \bar{d}^R & \bar{d}^B & \bar{d}^G & e^+ \\ \bar{u}^R & \bar{u}^B & \bar{u}^G & \bar{\nu} \end{array} \right)^i \end{equation} where $\alpha=1\ldots 4$ is an SU(4) index, $a,x=1,2$ are SU(2)$_{L,R}$ indices, and $i=1\ldots 3$ is a family index. The Higgs fields are contained in the following representations, \begin{equation} h_{a}^x=(1,\bar{2},2)= \left(\begin{array}{cc} {h_2}^+ & {h_1}^0 \\ {h_2}^0 & {h_1}^- \\ \end{array} \right) \label{h} \end{equation} (where $h_1$ and $h_2$ are the low energy Higgs superfields associated with the MSSM.) The two heavy Higgs representations are \begin{equation} {H}^{\alpha b}=(4,1,2)= \left(\begin{array}{cccc} u_H^R & u_H^B & u_H^G & \nu_H \\ d_H^R & d_H^B & d_H^G & e_H^- \end{array} \right) \label{H} \end{equation} and \begin{equation} {\bar{H}}_{\alpha x}=(\bar{4},1,\bar{2})= \left(\begin{array}{cccc} \bar{d}_H^R & \bar{d}_H^B & \bar{d}_H^G & e_H^+ \\ \bar{u}_H^R & \bar{u}_H^B & \bar{u}_H^G & \bar{\nu}_H \end{array} \right). \label{barH} \end{equation} The Higgs fields are assumed to develop VEVs, \begin{equation} <H>=<\nu_H>\sim M_{X}, \ \ <\bar{H}>=<\bar{\nu}_H>\sim M_{X} \label{HVEV} \end{equation} leading to the symmetry breaking at $M_{X}$ \begin{equation} \mbox{SU(4)}\otimes \mbox{SU(2)}_L \otimes \mbox{SU(2)}_R \longrightarrow \mbox{SU(3)}_C \otimes \mbox{SU(2)}_L \otimes \mbox{U(1)}_Y \label{422to321} \end{equation} in the usual notation. Under the symmetry breaking in Eq.\ref{422to321}, the Higgs field $h$ in Eq.\ref{h} splits into two Higgs doublets $h_1$, $h_2$ whose neutral components subsequently develop weak scale VEVs, \begin{equation} <h_1^0>=v_1, \ \ <h_2^0>=v_2 \label{vevs} \end{equation} with $\tan \beta \equiv v_2/v_1$. Below $M_{X}$ the part of the superpotential involving quark and charged lepton fields is just \begin{equation} W =\lambda^{ij}_UQ_i\bar{U}_jh_2+\lambda^{ij}_DQ_i\bar{D}_jh_1 +\lambda^{ij}_EL_i\bar{E}_jh_1+ \ldots \label{NMSSM} \end{equation} with the boundary conditions at $M_{X}$, \begin{equation} \lambda^{ij}_U=\lambda^{ij}_D=\lambda^{ij}_E. \label{boundary} \end{equation} The same Yukawa relations also occur in minimal $SO(10)$. \section{The Basic Strategy} Such Yukawa relations as in Eq.\ref{boundary} taken at face value are a phenomenological disaster. For example consider the minimal $SU(5)$ prediction $\lambda^{ij}_D=\lambda^{ij}_E$. After diagonalisation this leads to $\lambda_e = \lambda_d$, $\lambda_{\mu} = \lambda_s$, $\lambda_{\tau} = \lambda_b$, (at the scale $M_X$) and hence \begin{equation} \frac{\lambda_s}{\lambda_d} = \frac{\lambda_{\mu}}{\lambda_e} \end{equation} which is RG invariant and fails badly at low-energy. On the other hand the third family relation leads to the low-energy prediction (assuming the SUSY RG equations) $\lambda_b / \lambda_{\tau} \approx 2.4$ which works well. A possible fix is provided by the GJ texture, \begin{eqnarray} \lambda^E &=& \left(\begin{array}{ccc} 0 & \lambda_{12} & 0 \\ \lambda_{21} & 3 \lambda_{22} & 0 \\ 0 & 0 & \lambda_{33} \\ \end{array}\right) \\ \lambda^D &=& \left(\begin{array}{ccc} 0 & \lambda_{12} & 0 \\ \lambda_{21} & \lambda_{22} & 0 \\ 0 & 0 & \lambda_{33} \\ \end{array}\right), \end{eqnarray} With predictions (at $M_{X}$) \begin{eqnarray}&&\begin{array}{cc} \lambda_s = \lambda_\mu/3, & \lambda_d = 3\lambda_e, \\ \end{array} \end{eqnarray} which is viable. As it turns out, the idea of full top-bottom-tau Yukawa unification works rather well for the third family \cite{Yuk}, leading to the prediction of a large top quark mass $m_t>165$ GeV, and $\tan \beta \sim m_t/m_b$ where $m_b$ is the bottom quark mass. However Yukawa unification for the first two families is not successful, since it would lead to unacceptable mass relations amongst the lighter fermions, and zero mixing angles at $M_{X}$. In order to cure these problems, we require something akin to the GJ texture, in which the Yukawa relations are altered by group theoretical Clebsch coefficients, leading to enhanced predictivity. One interesting proposal has recently been put forward to account for the fermion masses in an SO(10) SUSY GUT with a single Higgs in the 10 representation \cite{Larry}. According this approach, only the third family is allowed to receive mass from the renormalisable operators in the superpotential. The remaining masses and mixings are generated from a minimal set of just three specially chosen non-renormalisable operators whose coefficients are suppressed by some large scale. Furthermore these operators are only allowed to contain adjoint 45 Higgs representations, chosen {}From a set of fields denoted $45_Y$, $45_{B-L}$, $45_{T_{3R}}$, $45_X$ whose VEVs point in the direction of the generators specified by the subscripts, in the notation of \cite{Larry}. This is precisely the strategy we wish to follow. We shall assume that only the third family receives its mass {}from a renormalisable Yukawa coupling. All the other renormalisable Yukawa couplings are set to zero. Then non-renormalisable operators are written down which will play the role of small effective Yukawa couplings. The effective Yukawa couplings are small because they originate {}from non-renormalisable operators which are suppressed by powers of the heavy scale $M$. We shall restrict ourselves to all possible non-renormalisable operators which can be constructed from different group theoretical contractions of the fields: \begin{equation} O_{ij}\sim (F_i\bar{F}_j )h\left(\frac{H\bar{H}}{M^2}\right)^n+{\mbox h.c.} \label{op} \end{equation} where we have used the fields $H,\bar{H}$ in Eqs.\ref{H},\ref{barH} and $M$ is the large scale $M>M_{X}$. The idea is that when $H, \bar{H}$ develop their VEVs such operators will become effective Yukawa couplings of the form $h F \bar{F}$ with a small coefficient of order $M_X^2/M^2$. Although we assume no intermediate symmetry breaking scale (i.e. SU(4)$\otimes $SU(2)$_L \otimes $SU(2)$_R$ is broken directly to the standard model at the scale $M_X$) we shall allow the possibility that there are different higher scales $M$ which are relevant in determining the operators. For example one particular contraction of the indices of the fields may be associated with one scale $M$, and a different contraction may be associated with a different scale $M'$. We shall either appeal to this kind of idea in order to account for the various hierarchies present in the Yukawa matrices, or to higher dimensional operators which are suppressed by a further factor of $M$. In the present model, although there are no adjoint representations, there will in general be non-renormalisable operators which closely resemble those in SO(10) involving adjoint fields. The simplest such operators correspond to $n=1$ in Eq.\ref{op}, with the $(H \bar{H})$ group indices contracted together. These operators are similar to those of ref.~\cite{Larry} but with $H\bar{H}$ playing the r\^{o}le of the adjoint Higgs representations. It is useful to define the following combinations of fields, corresponding to the different $H\bar{H}$ transformation properties under the gauge group in Eq.\ref{422}, \begin{eqnarray} (H \bar{H})_A & = & (1,1,1) \nonumber \\ (H \bar{H})_B & = & (1,1,3) \nonumber \\ (H \bar{H})_C & = & (15,1,1) \label{big} \\ (H \bar{H})_D & = & (15,1,3) \nonumber \end{eqnarray} It is straightforward to construct the operators of the form of Eq.\ref{op} explicitly, and hence deduce the effect of each operator. For example for $n=1$ the four operators are, respectively, \begin{equation} O^{A,B,C,D}_{ij}\sim F_i\bar{F}_jh\frac{(H \bar{H})_{A,B,C,D}}{M^2}+H.c. \label{n1} \end{equation} where we have suppressed gauge group indices. These operators lead to quark-lepton and isospin splittings, as shown explicitly below: \begin{eqnarray} O^{A}_{ij} & = & a_{ij} (Q_{i}\bar{U}_{j}h_2 + Q_{i}\bar{D}_{j}h_1 + L_{i}\bar{E}_{j}h_1 + H.c.) \nonumber \\ O^{B}_{ij} & = & b_{ij} (Q_{i}\bar{U}_{j}h_2 - Q_{i}\bar{D}_{j}h_1 - L_{i}\bar{E}_{j}h_1 + H.c.) \nonumber \\ O^{C}_{ij} & = & c_{ij} (Q_{i}\bar{U}_{j}h_2 + Q_{i}\bar{D}_{j}h_1 -3L_{i}\bar{E}_{j}h_1 + H.c.) \nonumber \\ O^{D}_{ij} & = & d_{ij} (Q_{i}\bar{U}_{j}h_2 - Q_{i}\bar{D}_{j}h_1 + 3L_{i}\bar{E}_{j}h_1 + H.c.) \label{effective} \end{eqnarray} where the coefficients of the operators $a_{ij},b_{ij},c_{ij},d_{ij}$ are all of order $\frac{M_X^2}{M^2}$. \section{Results and Conclusions} Using operators such as those above, together with more complicated $n=2$ operators, it is possible to account for the entire fermion mass spectrum. The successful ansatze \cite{422} involve 8 real parameters plus an unremovable phase. With these 8 parameters we can describe the 13 physical masses and mixing angles. Third family Yukawa unification leads to a prediction for $m_t (\mbox{pole}) = 130-200$ GeV and $\tan \beta = 35-65$, depending on $\alpha_S (M_Z)$ and $\bar{m}_b$. More accurate predictions could be obtained if the error on $\alpha_S(M_Z)$ and $\bar{m}_b$ were reduced. The analysis of the lower 2 by 2 block of the Yukawa matrices leads to 2 possible predictions for $\lambda_{\mu}/\lambda_s=3,4$ at the scale $M_X$ ($3$ is the GJ prediction). In the upper 2 by 2 block analysis we are led to 5 possible predictions for $\lambda_{d}/\lambda_e=2,8/3,3,4,16/3$. (again $3$ is the GJ prediction.) Finally, we have a prediction that $|V_{ub}|>0.004$\cite{422}. The high values of $\tan \beta$ required by our model (also predicted in SO(10)) can be arranged by a suitable choice of soft SUSY breaking parameters as discussed in ref.\cite{Carenaetal}, although this leads to a moderate fine tuning problem \cite{Yuk}. The high value of $\tan \beta$ is not stable under radiative corrections unless some other mechanism such as extra approximate symmetries are invoked. $m_t$ may have been overestimated, since for high $\tan \beta$, the equations for the running of the Yukawa couplings in the MSSM can get corrections of a significant size from Higgsino--stop and gluino--sbottom loops. The size of this effect depends upon the mass spectrum and may be as much as 30 GeV. For our results to be quantitatively correct, the sparticle corrections to $m_b$ must be small. This could happen in a scenario with non-universal soft parameters, for example. Not included in our analysis are threshold effects, at low or high energies. These could alter our results by several per cent and so it should be borne in mind that all of the mass predictions have a significant uncertainty in them. It is also unclear how reliable 3 loop perturbative QCD at 1 GeV is. Despite a slight lack of predictivity of the model compared to SO(10), the SU(4) $\otimes$SU(2)$_L \otimes$SU(2)$_R$ model has the twin advantages of having no doublet-triplet splitting problem, and containing no adjoint representations, making the model technically simpler to embed into a realistic string theory. Although both these problems can be addressed in the SO(10) model~\cite{new}, we find it encouraging that such problems do not arise in the first place in the SU(4)$\otimes$SU(2)$_L \otimes$SU(2)$_R$ model. Of course there are other models which also share these advantages such as flipped SU(5) or even the standard model. However, at the field theory level, such models do not lead to Yukawa unification, or have precise Clebsch relations between the operators describing the light fermion masses. It is the combination of all of the attractive features mentioned above which singles out the present model for serious consideration. \newpage
\section{Introduction} As QCD is the accepted theory of the strong interactions, it is no doubt desirable to understand all hadronic phenomena directly in terms of the fundamental fields of QCD. However, QCD being asymptotically free, perturbation theory is applicable only for very short distances and cannot cover the complete range of interest. At present lattice gauge simulations are the only way to study such systems. In a first approach, the static approximation is the natural choice, where the gluonic degrees of freedom are integrated out, and quark loops are ignored (the quenched approximation), giving rise to a potential between the stationary quarks. The potential of the quark-antiquark system, where this approach---leading to the familiar Wilson Loop---is very well known, has been calculated extensively in Monte Carlo lattice simulations. (For recent data, see e.g.~\cite{booth}.) The ground state potential of this static system has also been calculated in perturbation theory upto sixth order~\cite{fisch}. Here we shall describe how to generalize this procedure to multi-quark systems, especially to $(q\bar{q})^k$ systems. However, even with present-day computers, $q\bar{q}$ lattice simulations are still very demanding, and the amount of computations needed increases rapidly with the number of interacting quarks. Reliable models for multi-quark systems expressing their potentials e.g.\ in terms of the well known $q\bar{q}$-systems would therefore be of great help. Such two-body approximations have proven successful in many areas of physics, and these models can be formulated without difficulty. For the $qq\bar{q}\bar{q}$-system, which is the simplest one that can be considered consisting of two colour singlets, this model has been tested against numerical data from a Monte Carlo simulation \cite{paton}. For small distances the agreement has been found reasonable. It has also been observed \cite{mor} that the two-body model corresponds to lowest order perturbation theory. We shall be able to show that it is correct even to fourth order. To sixth order, however, three- and four-body forces begin to appear. \section{The Generalized Wilson Loop} While the concepts discussed below are of course well known in the context of the Wilson Loop for the $q\bar{q}$-system, we find it useful to start with rephrasing these concepts in the case of an arbitrary number of quarks, leading to a study of more complicated systems. When we have assembled a system of several quarks (and antiquarks), gluons will mediate a force between them. Treating this system in an approximation as a quantum mechanical system of several static quarks, the interactions between the quarks are incorporated into a potential. This assembly of quarks is then expected to propagate in time with the usual factor of $e^{-itH}$, where the interesting piece of the Hamilton operator $H$ is the potential energy. Thus, by calculating appropriate Green functions, the potentials of eigenstates of $H$ can be extracted. \subsection{Setting up Gauge Invariant States} \label{sin} Because of confinement, it makes sense only to talk about systems of quarks where the overall states have colour singlet quantum numbers. The problem with setting up say a $q\bar{q}$-system in a singlet is that the quark and antiquark are located a distance apart. This problem can be overcome by inserting the path ordered exponential $U(x,y,A) = {\cal P}e^{ig\int^x_y T^aA_a^{\mu}(z)dz_{\mu}}$ between the locations $x$ and $y$ of the quarks in the presence of the gauge potential $A$. Here $g$ denotes the coupling constant and $T^a$ the representation matrices. Thus $\bar{\psi}(x)U(x,y)\psi(y)|0\rangle$ will serve as a basis state in this case. We must also know how many basis states there are. When dealing with Green functions coming from Monte Carlo lattice simulations, they will have contributions from excited states of the gluonic field, and there are infinitely many of them even in the $q\bar{q}$-case. With suitable methods, the lowest potentials can be extracted, and several have been calculated for the quenched $q\bar{q}$-system -- see for example \cite{hunt}, \cite{mich}. The situation is different for Green functions calculated in perturbation theory. Here, unlike the lattice simulations, we can and must go to the infinite time limit. We do not expect to reach excited states of the gluon field in finite order perturbation theory, and thus the number of basis states for a system of several quarks is given by the usual arguments of group representation theory, e.g.\ one for the $q\bar{q}$-system and two for the $qq\bar{q}\bar{q}$-system. In the large time limit we expect that the effects of `introducing' the quarks into the vacuum will be irrelevant in comparison to their time evolution, and the notion of a potential makes sense. \subsection{Diagonalization} \label{diag} It may be shown that the state \[ |\mbox{ quarks }q_i\mbox{ and antiquarks }\bar{q}_j\mbox{ at time }t\rangle \] \begin{equation} = \bar{\psi}_{i}(-t,x_i)\ldots U(-t,x_i,y_i,A)\ldots \psi(-t,y_i)|0\rangle \label{e1} \end{equation} satisfies Schr\"odinger's equation. Forming the overlap of states at time $-t$ and $t$, we get an equation between Green functions and expressions of the form ${\bf A}_{ij}\stackrel{def}{=}\langle A_i|e^{-itH}|A_j\rangle$, where $|A_i\rangle$ stands for some basis state and we have introduced the matrix $\bf A$. By assuming a decomposition of these basis states into eigenstates of $H$, a diagonalization procedure will yield the potentials. In the case of the Green functions coming from lattice simulations, one considers a practical number of basis states, expands them in energy eigenstates and drops contributions with $e^{-itE_i}$ for energies $E_i$ above a certain limit. Of course we implicitly assume Wick-rotation. In perturbation theory, where a power expansion of $e^{-itE_i(g)}$ in the coupling $g$ will not be exponentially damped, we need to consider all linearly independent basis states, a number that is finite, as remarked in the last section. Because of this finiteness, we can find an invertible transformation to energy eigenstates, and the diagonalization is straightforward. In fact, given a matrix ${\bf A}$ satisfying certain consistency relations, we can perturbatively prove \cite{t} the existence of a time-independent basis transformation such that in this new basis ${\bf A}$ is not only diagonal, but its eigenvalues are of the form $e^{-itE_i(g)}$. Here the energy $E_i(g)$ of the $i$-th basis state, which can be calculated perturbatively, is for static quarks equal to the $i$-th potential (apart from an irrelevant constant, the rest mass). \subsection{Loops} \label{loops} What remains to be done is to bring the Green functions of the last paragraph to more familiar forms. Since we work within the static approximation, the full quark propagator in the presence of gauge fields can be calculated \cite{eichtenfeinberg}: \begin{eqnarray} S_0(x,y,A)&=&-i[{\cal P}e^{ig\int^x_yT^aA_a^{\mu}(z)dz_{\mu}}] e^{-im|x^0-y^0|}\delta(\vec{x}-\vec{y}) \times\nonumber\\ &&[\frac{1+\gamma^0}{2}\Theta(x^0-y^0)+\frac{1-\gamma^0}{2} \Theta(y^0-x^0)] \label{eq21} \end{eqnarray} We shall now outline how various contour integrations, i.e.\ loops arise. Considering the well-known $q\bar{q}$-case, we find a path-ordered line integral from antiquark to quark arising from the $U$ in eq.~(\ref{e1}), then the path-ordered line integral propagating the quark forward in time from eq.~(\ref{eq21}). Another $U$ and the antiquark propagating backwards in time close the rectangle of the familiar Wilson loop. Starting with the Green functions described below eq.~(\ref{e1}) and evaluating them for propagation from $-t/2$ to $t/2$, the following diagrammatic rule for calculating the Green function dealing with an arbitrary number $k/2$ of quark-antiquark pairs (i.e.\ $k$ quarks and antiquarks) partitioned into $q\bar{q}$ singlets is seen to hold: \begin{enumerate} \item Draw two horizontal lines, the lower denoting time $-t/2$, the upper $t/2$. Mark the position of every quark and antiquark on the lower line and once again vertically above it on the upper line. \item At the $-t/2$ level connect every quark-antiquark pair that is set up as a singlet at $-t/2$ with a line, having an arrow pointing from antiquark to quark. \item At the $t/2$ level connect every quark-antiquark pair that is set up as a singlet at $t/2$ with a line, the arrow in which points from quark to antiquark. \item Join the quarks at the $-t/2$ level with quarks at the same position at the $t/2$ level, arrow pointing upwards, i.e.\ forward in time. \item Join the antiquarks at the $t/2$ level with the antiquarks at the $-t/2$ level, arrow pointing downwards, i.e.\ backwards in time. \item Associate a path-ordered exponential of $e^{ig\oint_CT^aA_a^{\mu}(z)dz_{\mu}}$ together with a trace for every closed loop $C$ occurring. \item Determine the overall sign: If the pairings at the $-t/2$ level are the same as those on the $t/2$ level, there must be a $+$ sign. (This follows from the positivity of the norm on a Hilbert space if one lets $t\rightarrow 0$.) If this is not so, determine the sign of the permutation of antiquarks on the upper line that is necessary to give the same pairings as on the lower line. This is the overall sign. \item Multiply by $(\delta(\vec{0})e^{-imt})^{k}$, where $k$ is the total number of quarks and antiquarks. \item Insert the factor so obtained in the numerator\footnote{With $\eta$ we denote the ghost fields, with $\cal L$ the Lagrangian without fermions} of $\frac{\int[DA^a_{\mu}][D\eta_a^*][D\eta_a]e^{i\!\int\!d^4\!x[{\cal L}]}}{\int[DA^a_{\mu}][D\eta_a^*][D\eta_a]e^{i\!\int\!d^4\!x[{\cal L}]}}$ \end{enumerate} This gives the Green function in the chosen singlet structure. \section{The $qq\bar{q}\bar{q}$-Potentials} In $SU($N$)$ gauge theory with quarks in the fundamental representation, we want to calculate the $qq\bar{q}\bar{q}$-potential in perturbation theory to fourth order. It has been remarked in subsection~\ref{sin} that there are two independent basis states for this system, and one easily recognises a choice of these in the two possible ways of pairing the system into two quark-antiquark singlets. Assuming the first static quark at position $R_1$, the second at $R_2$, and the antiquarks at $R_3$ and $R_4$, we will label the two states $|A_1\rangle = {\bf 1}_{13}{\bf 1}_{24}$ and $|A_2\rangle = {\bf 1}_{14}{\bf 1}_{23}$. \subsection{Calculating the Green Functions} According to subsection~\ref{loops}, we encounter the following types of loops: \begin{center} \begin{picture}(350,150) \thicklines \put(0,35){\line(1,0){100}} \put(0,35){\vector(1,0){ 50}} \put(100,135){\line(-1,0){100}} \put(100,135){\vector(-1,0){50}} \put(300,135){\line(-1,0){100}} \put(300,135){\vector(-1,0){50}} \put(200,35){\line(1,0){100}} \put(200,35){\vector(1,0){50}} \put(0,135){\vector(0,-1){50}} \put(0,135){\line(0,-1){100}} \put(100,35){\vector(0,1){50}} \put(100,35){\line(0,1){100}} \put(200,135){\vector(0,-1){50}} \put(200,135){\line(0,-1){100}} \put(300,35){\line(0,1){100}} \put(300,35){\vector(0,1){50}} \put( 0, 15){$R_4$} \put(100, 15){$R_2$} \put(200, 15){$R_3$} \put(300, 15){$R_1$} \put(310, 35){$-t/2$} \put(310,125){$t/2$} \put(350,80){$C_{\langle A_1,-t/2|A_1,t/2\rangle}$} \end{picture} \end{center} and \begin{center} \begin{picture}(350,150) \thicklines \put(0,35){\line(1,0){300}} \put(0,35){\vector(1,0){150}} \put(100,135){\line(-1,0){100}} \put(100,135){\vector(-1,0){50}} \put(300,135){\line(-1,0){100}} \put(300,135){\vector(-1,0){50}} \put(200,40){\line(-1,0){100}} \put(200,40){\vector(-1,0){50}} \put(0,135){\vector(0,-1){50}} \put(0,135){\line(0,-1){100}} \put(100,40){\vector(0,1){50}} \put(100,40){\line(0,1){95}} \put(200,135){\vector(0,-1){50}} \put(200,135){\line(0,-1){95}} \put(300,35){\line(0,1){100}} \put(300,35){\vector(0,1){50}} \put( 0, 15){$R_4$} \put(100, 15){$R_2$} \put(200, 15){$R_3$} \put(300, 15){$R_1$} \put(310, 35){$-t/2$} \put(310,125){$t/2$} \put(350,80){$C_{\langle A_2,-t/2|A_1,t/2\rangle}$} \end{picture} \end{center} In calculating the Green functions, we will adopt dimensional regularization in dimension $D=4-2\epsilon$ with a mass scale $M$. A wave-function renormalization will remove infinities associated with a diagram of the form \input FEYNMAN \begin{center} \begin{picture}(10000,13000) \thicklines \put(5000,1500){\line(0,1){10000}} \put(5000,1500){\vector(0,1){5000}} \put(5000,4000){\circle*{500}} \put(5000,9000){\circle*{500}} \put(-2000,6500){${\tau}_1=$} \drawline\gluon[\E\REG](5200,9000)[1] \drawline\gluon[\SE\REG](\gluonbackx,\gluonbacky)[1] \drawline\gluon[\S\REG](\gluonbackx,\gluonbacky)[1] \drawline\gluon[\SW\REG](\gluonbackx,\gluonbacky)[1] \drawline\gluon[\W\REG](\gluonbackx,\gluonbacky)[1] \end{picture} \end{center} while a coupling constant renormalization is needed to make the sum \begin{center} \begin{picture}(10000,15000) \thicklines \put(5000,3500){\vector(0,1){10000}} \put(5000,8500){\circle*{500}} \put(5000,11000){\circle*{500}} \put(5000,6000){\circle*{500}} \put(5000,1500){${\tau}_2$} \put(5500,8400){$x$} \drawline\gluon[\E\FLIPPED](5100,6000)[1] \drawline\gluon[\NE\FLIPPED](\gluonbackx,\gluonbacky)[1] \drawline\gluon[\N\FLIPPED](\gluonbackx,\gluonbacky)[1] \drawline\gluon[\NW\FLIPPED](\gluonbackx,\gluonbacky)[1] \drawline\gluon[\W\FLIPPED](\gluonbackx,\gluonbacky)[1] \drawline\gluon[\W\FLIPPED](4900,8500)[2] \end{picture} \begin{picture}(10000,13000) \thicklines \put(10000, 3500){\line(0,1){10000}} \put(10000, 3500){\vector(0,1){ 5000}} \put(10000,8500){\circle*{500}} \put(10000,8500){\line(1,1){1200}} \put(10000,8500){\line(1,-1){1200}} \put(10000,8500){\line(-1,1){1200}} \put(10000,8500){\line(-1,-1){1200}} \drawline\gluon[\W\FLIPPED]( 9800,8500)[2] \end{picture} \end{center} finite. The full gluon propagator is renormalized in the usual way. Note that, contrary to a smallest distance regularization frequently encountered in this context, there is no renormalization of the quark mass $m$. Making the expansion in $\alpha = \frac{g^2{\it c_3}}{4\pi^2}$ rather than in $g$ and also expanding in time, a calculation yields for \begin{equation} {\bf A}={\bf A}\!(1) +\alpha{\bf A}\!(\alpha)+\alpha^2{\bf A}\!(\alpha^2)+\alpha t{\bf A}\!(\alpha t)+ \alpha^2t{\bf A}\!(\alpha^2t)+\frac{(\alpha t)^2}{2}{\bf A}\! (\frac{(\alpha t)^2}{2}) \label{eq36} \end{equation} the expressions: {\footnotesize \begin{eqnarray} \lefteqn{{\bf A}(1)= \left( \begin{array}{rr} N^2&-N\\-N&N^2 \end{array} \right)} \nonumber\\ &&\nonumber\\ \lefteqn{{\bf A}(\alpha)= \left(\!\! \begin{array}{ll} N^2 \left\{4 \ln(M^2R_{13}R_{24})\right\}& -N\left\{\ln\left(\frac{R^3_{13}R^3_{14}R^3_{23} R^3_{24}M^8}{R^2_{12}R^2_{34}} \right)-\mbox{e}_1-\mbox{e}_2\right\}\\ -N\left\{\ln\left(\frac{R^3_{13}R^3_{14}R^3_{23} R^3_{24}M^8}{R^2_{12}R^2_{34}} \right)-\mbox{e}_1-\mbox{e}_2\right\}& N^2 \left\{4 \ln(M^2R_{14}R_{23})\right\} \end{array} \!\!\right)} \nonumber\\ &&\nonumber\\ \lefteqn{{\bf A}(\alpha T)\,= \left( \begin{array}{ll} N^2\left\{-V_{s1}\right\}& -N\left\{-V_{s1}-V_{s2}+V_{d}\right\}\\ -N\left\{-V_{s1}-V_{s2}+V_{d}\right\}& N^2\left\{-V_{s2}\right\} \end{array} \right)} \nonumber\\ &&\nonumber\\ &&\nonumber\\ \lefteqn{{\bf A}(\alpha^2 T)\,=} \nonumber\\ &&\nonumber\\ && \left( \begin{array}{ll} \!\!N^2\left\{ -4V_{s1}\ln\left(M^2R_{13}R_{24}\right) \right. & -N \left\{ \left[-V_{s1}-V_{s2}+V_{d}\right]\times \right. \\ \left. \mbox{\ \ \ \ }+\frac{2{\it c_2}}{N{\it c_3}}\left[-V_{s2}+V_d\right]\times \right. & \left. \mbox{\ \ \ \ }\left[\ln\left(\frac{R^3_{13}R^3_{14}R^3_{23}R^3_{24}M^8} {R^2_{12}R^2_{34}}\right) -\mbox{e}_1-\mbox{e}_2\right] \right. \\ \left. \mbox{\ \ \ \ }\left[\ln\left(\frac{R_{14}R_{23}}{R_{12}R_{34}}\right)-\mbox{e}_1\right] \right\} & \left. \mbox{\ \ \ \ }+\frac{{\it c_1}}{2{\it c_3}}\left[\mbox{e}_1(V_d\!-\!V_{s1}) +\mbox{e}_2(V_d\!-\!V_{s2})\right]\!\! \right. \\ & \left. \mbox{\ \ \ \ }-\frac{{\it c_1}}{2{\it c_3}}\left[\ln(R_{12}R_{34})(V_{s1}\!+\!V_{s2}\!-\!2V_d)\right] \right.\!\! \\ & \left. \mbox{\ \ \ \ }+\frac{{\it c_1}}{2{\it c_3}}\left[\ln(R_{13}R_{24})(V_{s2}-V_d)\right] \right. \\ & \left. \mbox{\ \ \ \ }+\frac{{\it c_1}}{2{\it c_3}}\left[\ln(R_{14}R_{23})(V_{s1}-V_d)\right] \right\} \\ \\ \!\!-N\left\{ \left[-V_{s1}-V_{s2}+V_{d}\right]\times \right. & N^2\left\{ -4V_{s2}\ln\left(M^2R_{14}R_{23}\right) \right. \\ \left. \mbox{\ \ \ \ }\left[\ln\left(\frac{R^3_{13}R^3_{14}R^3_{23}R^3_{24}M^8} {R^2_{12}R^2_{34}} \right) -\mbox{e}_1 -\mbox{e}_2\right] \right. & \left. \mbox{\ \ \ \ }+\frac{2{\it c_2}}{N{\it c_3}}\left[-V_{s1}+V_d\right]\times \right. \\ \left. \mbox{\ \ \ \ }+\frac{{\it c_1}}{2{\it c_3}}\left[\mbox{e}_1(V_d-V_{s1}) +\mbox{e}_2(V_d-V_{s2})\right] \right. & \left. \mbox{\ \ \ \ }\left[\ln\left(\frac{R_{13}R_{24}}{R_{12}R_{34}}\right) -\mbox{e}_2\right] \right\} \\ \left. \mbox{\ \ \ \ }-\frac{{\it c_1}}{2{\it c_3}}\left[\ln(R_{12}R_{34})(V_{s1}\!+\!V_{s2}\!-\!2V_d)\right] \right. & \\ \left. \mbox{\ \ \ \ }+\frac{{\it c_1}}{2{\it c_3}}\left[\ln(R_{13}R_{24})(V_{s2}-V_d)\right] \right. & \\ \left. \mbox{\ \ \ \ }+\frac{{\it c_1}}{2{\it c_3}}\left[\ln(R_{14}R_{23})(V_{s1}-V_d)\right] \right\} & \end{array} \!\!\!\right) \nonumber\\ &&\nonumber\\ &&\nonumber\\ &&\nonumber\\ \lefteqn{{\bf A}\left(\frac{1}{2}(\alpha T)^2\right)\,=} \nonumber\\ &&\nonumber\\ && \left( \begin{array}{ll} N^2\left\{V_{s1}^2+\frac{{\it c_2}}{N{\it c_3}}(V_{s2}-V_d)^2 \right\} & -N \left\{ (V_{s1}+V_{s2}-V_d)^2 +\frac{{\it c_1}}{2{\it c_3}}\times \right. \\ \left. \mbox{\ \ \ \ } \right. & \left. \mbox{\ \ \ \ }\left[V_dV_{s1}+V_dV_{s2}-V_{s1}V_{s2}-V_d^2\right] \right\} \\ \\ -N\left\{ (V_{s1}+V_{s2}-V_d)^2 +\frac{{\it c_1}}{2{\it c_3}}\times \right. & N^2\left\{V_{s2}^2+\frac{{\it c_2}}{N{\it c_3}}\left(V_{s1}-V_d\right)^2 \right\} \\ \left. \mbox{\ \ \ \ }\left[V_dV_{s1}+V_dV_{s2}-V_{s1}V_{s2}-V_d^2\right] \right\} & \left. \mbox{\ \ \ \ } \right. \end{array} \right) \nonumber\end{eqnarray} } where, to save writing, the following shorthand notation has been adopted: \begin{eqnarray*} \mbox{edge}(\vec{R_3},\vec{R_1},\vec{R_4},\vec{R_2})\!\! &\!\stackrel{def}{=}\!& \!\!\int\!\!\int_{-1/2}^{1/2} \frac{(\vec{R_3}-\vec{R_1})(\vec{R_4}-\vec{R_2})\,\,\,dw\,dx} {\left[\frac{\vec{R_3}+\vec{R_1}}{2}\!+\!w(\vec{R_3}\!-\!\vec{R_1}) \!-\!\frac{\vec{R_4}\!+\!\vec{R_2}}{2}\!-\!x(\vec{R_4}\! -\!\vec{R_2}) \right]^2}\\ \mbox{e}_1&\stackrel{def}{=}&\mbox{edge}(R_1,R_3,R_2,R_4)\\ \mbox{e}_2&\stackrel{def}{=}&\mbox{edge}(R_1,R_4,R_2,R_3)\\ V_{s1}&\stackrel{def}{=}&V(R_{13})+V(R_{24})\\ V_{s2}&\stackrel{def}{=}&V(R_{14})+V(R_{23})\\ V_{d}&\stackrel{def}{=}&V(R_{12})+V(R_{34})\\ \end{eqnarray*} \vspace{-5 ex} \[ \begin{array}{rrrrr} {\it c_1}\delta_{ab}&=&\sum_{cd}{\it f}_{acd}{\it f}_{bcd}&=&N\delta_{ab} \\ {\it c_2}\delta_{ab}&=&{\it Tr\/}[T^aT^b]&=&\frac{1}{2}\delta_{ab} \\ {\it c_3}{\bf 1}&=&\sum_a T^aT^a&=&\frac{N^2-1}{2N}{\bf 1} \end{array} \] and $V(R_{pq})$ is the two-body potential between a quark `p' and an antiquark `q' a distance $R_{pq}$ apart. Diagonalization yields for the two possible energy eigenstates the two potentials correct to fourth order: \begin{eqnarray} V_0\!&=&\!\frac{ \left(N^2\!-\!2\right) \left(V_{s1}\!+\!V_{s2}\right) +2V_{d} -N\sqrt{ N^2\left(V_{s1}\!-\!V_{s2}\right)^2 +4\left(V_{s1}\!-\!V_d\right)\!\left(V_{s2}\! -\!V_d\right) } }{2\left(N^2-1\right)} \nonumber\\ V_1\!&=&\!\frac{ \left(N^2\!-\!2\right) \left(V_{s1}\!+\!V_{s2}\right) +2V_{d} +N\sqrt{ N^2\left(V_{s1}\!-\!V_{s2}\right)^2 +4\left(V_{s1}\!-\!V_d\right)\!\left(V_{s2}\! -\!V_d\right) } }{2\left(N^2-1\right)} \nonumber\\ \label{potentials} \end{eqnarray} These potentials are exactly equal to those given by a naive two-body model (see~\cite{paton})---and are one of the main conclusions of this work. \subsection{Consistency Relations and Edge Effects} When integrating out the gluonic degrees of freedom and reducing a field theory to a quantum mechanical system, we have made assumptions which may need some justification. This is especially so in the perturbative case where we have postulated a finite number of energy eigenstates. We, therefore, want to have a look at the consistency of this approach. There are certain relations which must hold in order to guarantee proper exponentiation. We have remarked at the end of subsection~\ref{diag} that after diagonalization we expect the diagonal entries of the matrix $\bf A$ to be (apart from a normalization factor) of the form $e^{-itE_i(g)}$. An actual calculation will give a power expansion in $g$ and $t$ for these diagonal entries. This will not only determine the energy $E_i(g)$ as the coefficient of $t$, but will also allow us to check the consistency of our calculation by inspecting the relations between the coefficients of higher powers of $t$. We were able to verify these consistency relations in our calculation to fourth order. Another important point to notice is that the terms we have abbreviated $\mbox{edge}(\vec{R_p},\vec{R_q},\vec{R_r},\vec{R_s})$ have cancelled in eq.~(\ref{potentials}). These terms are coming from the space-like contour integrations. In subsection~\ref{sin} we had introduced these space-like contour integrations via the path-ordered exponential $U$ in order to establish colour singlets. The exact integration contour in $U$ was of course arbitrary, and the physical potentials must not depend on it. Another way of looking at it is that in the large time limit the processes associated with bringing the quarks into their position in the distant past (and removing them again in the distant future) should become irrelevant. One may also make an adiabatic argument, turning on the coupling $g$ in the distant past and off again in the distant future, thus making the notion of a colour singlet state at these times well-defined in spite of the spatial separation of the quarks. In any case, for the potentials of eq.~(\ref{potentials}) to be meaningful, they must not involve terms originating in contour integrations in the distant past or future, and indeed they do not. \subsection{Three- and Four-Body Forces in Sixth Order} The fact that a straightforward two-body model is correct also to next-to-leading order may be surprising in light of the non-abelian nature of QCD. Hence we want to mention that we believe that the two-body model fails at sixth order as three- and four-body forces appear at this order. While many diagrams that seem to give rise to such deviations from the two-body model actually vanish, we see no reason why for example effects from the following two diagrams should cancel in the calculation of the four-quark potentials to sixth order. \begin{center} \begin{picture}(10000,13000) \thicklines \put(0,11500){\vector(0,-1){5000}} \put(10000,1500){\vector(0,1){5000}} \put(0,1500){\vector(1,0){5000}} \put(10000,11500){\vector(-1,0){5000}} \put(0,11500){\line(0,-1){10000}} \put(10000,1500){\line(0,1){10000}} \put(0,1500){\line(1,0){10000}} \put(10000,11500){\line(-1,0){10000}} \put(0,9000){\circle*{500}} \put(10000,4000){\circle*{500}} \drawline\gluon[ \E \REG]( 250, 9000)[12] \drawline\gluon[\SE \REG](\gluonbackx,\gluonbacky)[1] \drawline\gluon[ \E\FLIPPED]( 9800, 4000)[3] \drawline\gluon[\NE\FLIPPED](\gluonbackx,\gluonbacky)[1] \end{picture} \begin{picture}(10000,13000) \put(5000,6500){\circle{1000}} \end{picture} \begin{picture}(10000,13000) \thicklines \put(0,11500){\vector(0,-1){5000}} \put(10000,1500){\vector(0,1){5000}} \put(0,1500){\vector(1,0){5000}} \put(10000,11500){\vector(-1,0){5000}} \put(0,11500){\line(0,-1){10000}} \put(10000,1500){\line(0,1){10000}} \put(0,1500){\line(1,0){10000}} \put(10000,11500){\line(-1,0){10000}} \put(0,9000){\circle*{500}} \put(0,4000){\circle*{500}} \drawline\gluon[ \W\FLIPPED]( 100, 9000)[3] \drawline\gluon[\SW\FLIPPED](\gluonbackx,\gluonbacky)[1] \drawline\gluon[ \W \REG]( 100, 4000)[3] \drawline\gluon[\NW \REG](\gluonbackx,\gluonbacky)[1] \end{picture} \end{center} \begin{center} \begin{picture}(10000,15000) \thicklines \put(0,13500){\vector(0,-1){5000}} \put(10000,3500){\vector(0,1){5000}} \put(0,3500){\vector(1,0){5000}} \put(10000,13500){\vector(-1,0){5000}} \put(0,13500){\line(0,-1){10000}} \put(10000,3500){\line(0,1){10000}} \put(0,3500){\line(1,0){10000}} \put(10000,13500){\line(-1,0){10000}} \put(0,11000){\circle*{500}} \put(10000,6000){\circle*{500}} \drawline\gluon[ \E \REG]( 250,11000)[12] \drawline\gluon[\SE \REG](\gluonbackx,\gluonbacky)[1] \drawline\gluon[ \E\FLIPPED]( 9800, 6000)[3] \drawline\gluon[\NE\FLIPPED](\gluonbackx,\gluonbacky)[1] \end{picture} \begin{picture}(10000,15000) \put(5000,8500){\circle{1000}} \end{picture} \begin{picture}(10000,15000) \thicklines \put(0,13500){\vector(0,-1){5000}} \put(10000,3500){\vector(0,1){5000}} \put(0,3500){\vector(1,0){5000}} \put(10000,13500){\vector(-1,0){5000}} \put(0,13500){\line(0,-1){10000}} \put(10000,3500){\line(0,1){10000}} \put(0,3500){\line(1,0){10000}} \put(10000,13500){\line(-1,0){10000}} \put(0,11000){\circle*{500}} \put(10000,6000){\circle*{500}} \drawline\gluon[ \W\FLIPPED]( 0,11000)[3] \drawline\gluon[\SW\FLIPPED](\gluonbackx,\gluonbacky)[1] \drawline\gluon[ \W \REG](10250, 6000)[13] \drawline\gluon[\NW \REG](\gluonbackx,\gluonbacky)[1] \end{picture} \end{center} \begin{center} \begin{picture}( 3000, 8000) \put( 0, 4000){where} \end{picture} \begin{picture}( 5000, 8000) \put( 2500, 4000){\circle{1000}} \end{picture} \begin{picture}( 5000, 8000) \put( 0, 4000){stands for} \end{picture} \begin{picture}(15000,8000) \thicklines \put(5000,4000){\circle*{500}} \put(10000,4000){\circle*{500}} \drawline\gluon[\E \REG ]( 5250, 4000)[4] \drawline\gluon[\NW\FLIPPED]( 5000, 4000)[2] \drawline\gluon[\SW\FLIPPED]( 5000, 4000)[2] \drawline\gluon[\NE\REG ](10000, 4000)[2] \drawline\gluon[\SE\REG ](10000, 4000)[2] \end{picture} \end{center} Hence three- and four-body forces will be introduced. In general, their nature seems to be complicated, but for some geometries simplifications are possible; e.g. for the four quarks on the corners of a regular tetrahedron there will be no contribution from quark self- interactions to four-body forces to sixth order. \section{Relation of our Result to Lattice Simulations} Looking at the Monte Carlo lattice calculations for the $qq\bar{q}\bar{q}$-system discussed in~\cite{paton}, we observe that for small interquark distances of a few lattice spacings (with a lattice spacing of $a\approx 0.12$ fm) the two-body model gives a reasonable approximation in the sense that the four-quark potentials calculated from eq.~(\ref{potentials}) using the Monte Carlo two-body potentials are comparable to the four-quark potentials from the lattice simulation. The agreement improves the smaller the distances get. By comparing the perturbative (i.e.\ $1/R$) and non-perturbative (i.e. linear) part in the usual parametrization of the $q\bar{q}$-potential, one would expect to start entering the perturbative regime at distances of about two lattice spacings. However, at that stage the approximation provided by the two-body model is already very good. The fact that the two-body model is correct to fourth order in perturbation theory certainly suggests that it should be a reasonable approximation in the perturbative domain. So our result supports the belief that the results of the lattice simulations for small enough distances indeed are correlated to continuum perturbation theory, and thus that continuum physics is extracted from the Monte Carlo calculations. \vspace{5 ex} \paragraph{Acknowledgements} One of the authors (J. L.) wants to express his gratitude to the `Stiftung Maximilianeum' which made possible his stay at Oxford, during which time most of this work was carried out.
\section{Introduction} \setcounter{equation}{0} The machinations of black holes have been studied extensively in recent years using two dimensional models \cite{CGHS,RST}. In\cite{BGHS,SHE},the original {\sl CGHS} model was solved numerically for explicitly static equilibrium scenarii. The {\sl CGHS} lagrangian is a dilaton gravity model which is made up of a classical part, which comes directly from low energy string theory in two dimensions, and a quantum correction described by Polyakov in\cite{POL} which takes into account the one-loop effects of the matter fields. The number of these fields can be proliferated so that the effect of other quantum corrections is small compared to that of the matter. In another paper, by Lemos and S\` a\cite{LEMSA}, the classical lagrangian considered is more general than that of the {\sl CGHS} model. A variable multiplicative parameter is included in the kinetic dilaton term. By choosing certain values for this parameter, a set of classical models which includes low energy string theory in two dimensions, Jackiw-Teitelboim theory, and planar general relativity is obtained. In this paper, the idea is to combine and extend the work of \cite{LEMSA} and \cite{BGHS,SHE}. The more general classical core lagrangian of \cite{LEMSA} will be combined with the Polyakov quantum matter correction, and a set of static numerical solutions to various models with and without the correction are displayed. The static black holes in equilibrium with Hawking radiation can be studied numerically. One motivation for studying such solutions is to understand the `quantum' singularity which was discovered shortly after the appearance of the original {\sl CGHS} paper\cite{RSTA,BAN}. Birnir {\it et al} noted that this singularity occurred at the finite dilaton value, and that the metric was actually finite there, unlike the classical case. We would like to investigate this further here. The static solutions might be a candidate final state of black hole evaporation, {\it i.e.} as massive remnants. This was rejected in\cite{BGHS} since the ADM mass for these solutions is divergent because there is non-zero radiation density out to infinity. Here we shall find the expression for the ADM mass in that case and show that it is indeed infinite. For equilibrium in two dimensions, this divergence is actually necessary, but we shall see that the solutions are nevertheless interesting. In the following section, a general two dimensional homogeneous dilaton gravity model is introduced, whose field equations for static solutions are written down. The initial conditions for regular-horizon spacetimes are then given. In section three, a general introduction to the results is given. Static regular horizon solutions in which the dilaton is initially and remains sub- or super-critical are then found for a range of classical cores with and without quantum corrections. The solutions being static, and numerical, it is difficult to be precise about global structure away from the singularities at infinity, though one can give some details about the singularity itself and the horizon. We restrict a fuller description and interpretation to the case whose classical core is that of planar general relativity. In section four are the conclusion and discussion. \section{General Homogeneous Two-dimensional Model} \setcounter{equation}{0} \label{sect:models} \subsection{Introduction} A general homogeneous lagrangian with semi-classical minimal scalars is of the form \begin{equation} I = \frac 1 {2\pi}\int d^2 x \sqrt{\pm g} \Big[ R\tilde\chi(\Phi)+4\omega e^{-2\phi}{\nabla \phi}^2+V(\Phi) -\frac {\kappa} 4 {\nabla Z}^2-\frac 1 2 \sum_{i=1}^N(\nabla f_i)^2 \Big] -\frac 1 {\pi} \int d \Sigma \sqrt{\pm h} K\tilde\chi, \label{eq:gen} \end{equation} where $\tilde\chi=\frac {\kappa} 2 Z + \Phi,$ and $\Phi$ is a function of the dilaton field $\phi$ and any other fields that are required. The terms involving Z in the volume term of this action have replaced the usual Polyakov term\cite{POL} $\frac {\kappa} 4 R(x) \int d^2 x' \sqrt{-g(x')} G(x,x')R(x') $ which comes from the matter contribution to the associated path integral. $G(x,x')$ is the scalar Green's function. The trace anomaly of the Z scalar field is that of the N minimal scalars. One could choose the function of the dilaton field so as to make the theory have vanishing central charge\cite{ABC}, but for simplicity models for which $V(\Phi)=4\lambda^2,$ will be restricted to here. The function $\tilde\chi$ takes into account both the classical coupling of the dilaton to gravity, and any one-loop terms which come from quantising additional fields. The classical part will be taken to be $e^{-2\phi},$ and the one-loop corrections to be those of the {\sl CGHS} model, so that $\tilde\chi=e^{-2\phi}-\frac {\kappa} 2 (\epsilon\phi-Z)$, where $\epsilon=0$ \footnote{The {\sl RST} \cite{RST} model has $\epsilon=1,\omega=1$.}. The {\sl CGHS} model is regained when $\omega=1.$ If one tries to work in the two-dimensional analogy of the `Eddington-Finkelstein' gauge\cite{JH2}, the action and field equations are still complicated although the work on entropy which depends on the position of the horizon might be simplified. However for static solutions one should use the conformal gauge where the line element is \begin{equation} dl^2=-e^{2\rho}dx^{+}dx^{-}. \label{eq:conf1} \end{equation} The field equations then imply that $Z=2\rho$, up to a solution of the wave equation. The action (\ref{eq:gen}) now becomes \begin{equation} I=-\frac 1 {\pi} \int d^2 x \sqrt{g} \Big[e^{-2\phi} (2\partial_- \partial_+ \rho -4\omega\partial_- \phi\partial_+ \phi+ \lambda^2e^{2\rho} )+ \frac 1 2 \partial_+ f_{i}\partial_- f_{i} -\kappa(\partial_- \rho\partial_+ \rho + \epsilon\phi\partial_- \partial_+ \rho) \Big] \label{eq:action} \end{equation} The surface term is omitted from now on. \subsection{Field Equations} \label{sect:fe} The following applies to static solutions which are functions of the static variable $s=-x^+x^-$ only. Terms in $f_i$ have been set to zero. Denoting $'=\frac d {ds}$, the field equations become \begin{equation} Q(\phi) [\phi '' +\frac 1 s \phi '] = 2{\phi '}^2(P(\epsilon,\phi)-\frac 1 2 \omega\kappa e^{2\phi})- \frac {\lambda^2} {2s}e^{2\rho}(P(\epsilon,\phi)-\frac 1 2 \kappa e^{2\phi}) \label{eq:field1} \end{equation} \begin{equation} Q(\phi) [\rho '' +\frac 1 s \rho '] = 2\omega{\phi '}^2(2-P(\epsilon,\phi))-\frac {\lambda^2} {2s}e^{2\rho}(2\omega-P(\epsilon,\phi)) \label{eq:field2} \end{equation} where $P(\epsilon,\phi)=1+\frac {\epsilon\kappa} 4 e^{2\phi}$ and $Q(\phi)=P(\epsilon,\phi)^2-\omega\kappa e^{2\phi}$. The constraint equations are \begin{equation} P(\epsilon,\phi)(\phi '' -2\rho '\phi ') +2(\omega-1){\phi '}^2 + \frac 1 2 \kappa e^{2\phi} ({\rho '}^2 -\rho ''+ \frac t {s^2})=0 \label{eq:con} \end{equation} where t is given by the boundary conditions required. These equations which will reduce in the case of $\omega=1$ to either the CGHS ($\epsilon=0$) or RST ($\epsilon=1$) model. \subsection{Initial Conditions} One can solve the dynamical equations numerically for $\rho,$ $\phi$ and their derivatives, by rewriting them as a coupled set of four first order differential equations. The boundary conditions chosen are such that the origin in $s$ is regular. This requires that \begin{equation} s\rho''(0)=s\phi''(0)=0, \label{eq:cdn1} \end{equation} \begin{equation} t=0, \label{eq:cdn2} \end{equation} A shift in $\rho$ allows one to remove $\lambda$ from the equations. $\phi$ can also be redefined so that the equations are independent of $\kappa$. One then finds that the derivatives at the origin should be \begin{equation} \rho '(0)=\frac {-\frac 1 2 e^{2\rho_{0}}[P(\phi_0)-2\omega]} {Q(\phi_0)} \label{eq:dero} \end{equation} \begin{equation} \phi '(0)=\frac {-\frac 1 2 e^{2\rho_{0}}[P(\phi_0)-\frac 1 2 e^{2\phi_{0}}]} {Q(\phi_0)} \label{eq:derphi} \end{equation} These equations reduce to those of \cite{BGHS} in the case $\omega=1, \epsilon=0.$ Varying the initial value of $\rho$ simply scales the equations. The initial value of $\phi$ at the origin is related to the `size' of the black hole. That is, moving towards the critical value initially, reduces the coordinate distance to the singularity from the horizon. \subsection{Choice of $\omega$-Parameter} \label{sect:which} In order to choose the smallest set of values of $\omega$ each of which produce different behaviour, one can consider the expression (\ref{eq:rq}) for the curvature, $R=-8e^{-2\rho}\frac {d} {ds}+s\frac {d^2} {ds^2}\rho$. the field equations (\ref{eq:field1},\ref{eq:field2}) and the values of $\omega$ and $\phi_0$ which change the sign of the initial values of $\frac {d\rho} {ds}$ and $\frac {d\phi } {ds}$. The critical value of $\phi$ \footnote{$\omega$ is multiplied by $\kappa$ in the logarithm if $\kappa$ has not been scaled out.} is given by \begin {equation} \phi_{cr}=-\frac 1 2 \log\omega. \end{equation} Singularities will therefore occur at increasingly weak coupling as $\omega$ is increased. The CGHS model(see section \ref{sect:models}.1) corrections are used in the following, {\it i.e.} $\epsilon=0$ and so $P(\phi,\epsilon)=1$. The initial value of $\frac {d\phi } {ds}$ is then zero at $\phi_0=\frac 1 2 \log 2$, unless $\omega=\frac 1 2 $. One can divide the cases first into those for which $\omega>\frac 1 2 $, $\omega=\frac 1 2 $, and $\omega<\frac 1 2 $. Then for the former case one has $\phi_0<\phi_{cr}$,$\phi_{cr}<\phi_0<\frac 1 2 \log 2$ and $\phi_0>\frac 1 2 \log 2$. When $\omega<\frac 1 2 $, we have $\phi _0<\frac 1 2 \log 2 $, $\frac 1 2 \log 2<\phi_0<\phi_{cr}$ and $\phi_0>\phi_{cr}$. For $\omega=\frac 1 2$, one simply has sub- and super-critical initial values, $\phi_0<\phi_{cr}$ and $\phi_0>\phi_{cr}$. One can also compare with the classical counterparts for which $\kappa=0$ in the field equations. The critical values of $\phi$ do not exist in the classical case. By inspection of the field equations (\ref{eq:field1},\ref{eq:field2}), the regions from which one would like to consider a value of $\omega$ are \begin{equation} {\omega<0; \omega=0; 0<\omega<\frac 1 2 ; \omega=\frac 1 2 ; \frac 1 2 <\omega<1; \omega=1; \omega>1} \end{equation} Lemos and Sa\cite{LEMSA} also show that the global structure differs for the cases $1<\omega<2$,$\omega=2$,and $\omega>2$. The numerical analysis does not distinguish qualitatively between these cases. We leave until later a fuller discussion of the case $\omega=\frac 1 2$. \newpage \section{Solutions} Since the corrected solutions are static by ansatz, and there is in general non-zero radiation density outside the black hole due to the Polyakov term associated with the minimal matter fields, they represent equilibrium scenarii. The ADM mass may be calculated as follows\cite{W}. Let $g_{ab}=\eta_{ab}+h_{ab}$ , and $\phi=\Phi_{L}+\varphi$, be perturbations from flat space $\eta_{ab}$, and from linear dilaton $\Phi_{L}$, where $h_{ab}$ and $\varphi$ vanish at infinity. The total mass measured by an observer at right infinity is given by \begin{equation} M = \int t_{0\mu}\xi^{\mu} dx \label{eq:mass} \end{equation} where $t_{0j}$ comes from the linearised energy-momentum tensor for the classical theory, $\xi^j$ is a timelike Killing vector, and $x$ is a suitable radial coordinate. For this calculation, one needs the generalised asymptotic expansions of $\rho$ and $\phi$. In the case $\omega=1$, we have these expressions\cite{SHE}. Below, we shall note the result for this case, which is representative of the $\omega>0$ cases. It would seem that $M\to\pm\infty$ except when $\omega=0$. This is due to the thermodynamics peculiar to two dimensions. There is by construction, an horizon on $s=0$ in all the cases. For example, in the case $\omega=1$, when one reproduces the classical black hole of Witten\cite{W}, there is a curvature singularity at finite negative $s$, behind the horizon at the origin. One can see how the distance from the origin to the singularity decreases as one goes toward the critical value of dilaton. This corresponds to a smaller black hole, which would appear later in a sequence of static black holes that one might use to represent black hole evolution. However, the sequence can never be complete because of the divergences as one approaches the critical value. \subsection{Classical Solutions} There will be given a set of plots of the numerical regular-horizon solutions of the model with the various values of the parameter $\omega$ considered in \cite{LEMSA} for the classical case which has $\kappa=0$. For $\kappa=0$, one can show that there exists a timelike killing vector, so the most general solution is static \cite{GIBPERB}. In the classical case, the initial value of the gradient of $\phi$ is given by \begin{equation} \phi '(0)=-\frac 1 2 e^{2\rho_{0}} \label{eq:derphii} \end{equation} and $\rho_{0}=0$ is chosen in each case. This initial value is independent of $\omega$. For the initial gradient of $\rho$ \begin{equation} \rho '(0)=-\frac 1 2 (2\omega-1) e^{2\rho_{0}} \label{eq:derrhoi} \end{equation} which clearly depends on $\omega$ and goes through zero at $\omega=\frac 1 2 $. Let the operator \begin {equation} D=\frac {d} {ds}+s\frac {d^2} {ds^2}. \label{eq:D} \end{equation} The classical equations are \begin{equation} D \phi = 2s{\phi '}^2-\frac 1 2 e^{2\rho}, \label{eq:field1c} \end{equation} \begin{equation} D \rho = 2s\omega{\phi '}^2-\frac 1 2 e^{2\rho}(2\omega-1). \label{eq:field2c} \end{equation} Notice that $\phi$ only appears in the equations as a derivative of s. This means that it will not matter as far as qualitative changes are concerned what the initial value of $\phi$ is: there are no {\it critical} values of $\phi_0$. The initial conditions given above are applied, which ensure that the solution is regular at the horizon, $s=0$. The initial value of $\rho$ simply scales, and is taken in every case to be zero. The coordinates $x^+,x^-$ are analogous to the Kruskal coordinates in the Schwarzschild solution, and cover the extended manifold. The two coordinate invariant functions $\phi$ and the curvature scalar R, are plotted, along with the metrical factor $\rho$ for several values of $\omega$. In order to obtain solutions which are regular at the origin, one has to integrate from the origin in both directions using particular initial conditions on $\rho$ and $\phi$ and their derivatives. The key points are to note singularities, or lack of singularities in the curvature and whether they occur at weak or stong coupling in $\phi$, and also to note divergences in $\rho$ and/or $\phi$, whilst the curvature is finite. Further physical conclusions are difficult to make since these are numerical solutions which are static, and thus effectively one-dimensional. The following gives the legend for the numerical solution plots: \vbox{\[ \begin{array}{c} \epsfxsize=0.35\displaywidth \epsfbox{legend.ps}\\ \end{array}\]} Our classical numerical results are in agreement with the analytical solutions of \cite{LEMSA} for their $A>|B|$, with $\lambda^2>0$. It is necessary to reproduce these so that one can compare with the new semi-classical solutions given later. In the following, brief physical comments are made upon each of the solutions, which should be read in conjunction with FIGS. 1-3. \vspace{0.3in} \centerline{\bf A. \ $\omega<0$} \smallskip We take for example $\omega=-1$. Towards positive $s$, $\phi$ diverges to minus infinity, whilst $\rho\to\infty$, at finite coordinate distance. The curvature is approximately constant and negative. Thus we have a timelike right infinity, at weak coupling. To the left we have $\phi\to\infty$, whilst $\rho\to-\infty$. The curvature goes to $-\infty$ so we have a timelike singularity. The extended manifold is given in \cite{LEMSA}, as for all the following classical cases. \smallskip \vbox{ \[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{cpfig.ps}\\ \end{array}\] FIG. 1. The dilaton field for the five cases of $\omega$-parameter at the classical level.} \vspace{0.3in} \centerline{\bf B. \ $\omega=0$ } This is the Jackiw-Teitelboim theory considered in \cite{JT}. It represents constant curvature anti-de Sitter space with strong coupling to the left, and weak coupling to the right. The second field equation above (\ref{eq:field2c}) is precisely the statement that the curvature scalar will be $R=-4$, and this is borne out in FIG. 3. \vspace{0.3in} \centerline{\bf C. \ $\omega=\frac 1 2 $ } \smallskip This is planar general relativity. The gradient of $\rho$ is initially zero. In the above cases it is positive, and becomes increasingly negative as $\omega$ increases. There is a spacelike singularity at strong coupling($\phi\to\infty$) to the left as in the previous case and for all $\omega>0$. To the right, $\phi$ diverges to minus infinity, and the curvature tends to a constant negative value. There is thus a black hole with a timelike right infinity. \vbox{\[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{crhfig.ps}\\ \end{array}\] FIG. 2. The conformal factor for the five cases of $\omega$-parameter at the classical level.} \vspace{0.3in} \centerline{\bf D. \ $\omega=1$ } \smallskip This is the classical black hole of Witten\cite{W}, which was found in low energy string theory. There is a spacelike singularity at strong coupling($\phi\to\infty$), and the curvature tends to zero at right infinity, which is null. \vspace{0.3in} \centerline{\bf E. \ $\omega>1$} \smallskip We consider $\omega=\frac 3 2$. This has the spacelike singularity at left infinity at strong coupling($\phi\to\infty$), but at right infinity, $\phi$, $\rho$ and the curvature all go to minus infinity logarithmically. The rate at which the curvature does so increases as $\omega$ increases. This means that these spacetimes have singularities to the right and left, spacelike and timelike respectively. More details can be found in \cite{LEMSA}. \vbox{ \[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{crfig.ps}\\ \end{array}\] FIG. 3. The scalar curvature at the classical level.} \subsection{Quantum Corrections} It is clear that the quantum corrections make significant qualitative changes to the overall structure of the spacetime, as can be seen by comparing the $\phi$, $\rho$, and curvature scalar plots for classical and quantum cases. But one would expect such differences from looking for example at the expressions for the curvature scalar as a function of $\omega$, $\phi$, and $\rho$. The curvature scalar is given by \begin {equation} R=-8e^{-2\rho}D\rho. \end{equation} The field equations can be used to rewrite this expression as \begin {equation} R_{cl}=4(2\omega-1)-16\omega s{\phi'}^2 e^{-2\rho}, \end{equation} \begin {equation} R_{q}=\frac {4(2\omega-1)-16\omega s{\phi'}^2 e^{-2\rho}} {P(\epsilon,\phi)^2- \omega e^{2\phi}} \label{eq:rq} \end{equation} in the classical and quantum-corrected cases respectively. Therefore at weak coupling $(\phi<<0)$, the two quantities may be approximately equal, but clearly, there are large effects near the critical value. Indeed, several examples will be seen of the `semi-classical' type of singularity which happens when $\phi$ hits the critical value. Let us define the {\it semi-classical singularity} to be one where the dilaton field is finite. The coordinates may or may not diverge at this point, but this clearly depends on the coordinate system. The key difference between this singularity and the classical ones is simply that the dilaton field no longer diverges there. Note that if $\omega\leq 0$ there can be no semi-classical singularities, although in some cases there are still qualitative differences in causal structure due to the corrections. \vbox{ \[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{q1pfig.ps}\\ \end{array}\] FIG. 4. The dilaton at the semi-classical level for subcritical initial value.} \subsection{Quantum Solutions} In the work of \cite{BGHS}, given in their FIG. 2, $\phi_{cr}=-2$, and $\kappa=e^4$. One should note again that this singularity occurs at finite $\phi$ and $\rho$. Indeed for this choice of $\kappa$, the singularity is actually at weak coupling $g_s=e^{-2}$. The same equations have been solved here in the case $\omega=1$, with the same initial conditions but different $\kappa$ and initial value of dilaton. The results are qualitatively identical as expected. In the new plots given here, $\kappa$ and $\lambda$ have been scaled out of the equations by using field redefinitions of the variables $\rho$ and $\phi$. For the quantum case, in general, three sets of graphs for each value of $\omega$ are needed, whereas in all the classical plots, $\phi$ ranges $(-\infty,+\infty)$. In FIG. 4, for $\omega=1$, however, $\phi<0$. Witten\cite{W}, regarded the dilaton field as a coordinate-independent measure of an observers position. Viewing it as such, one might believe some of space to have been omitted as the region $\phi\ge 0$ does not exist. This is why we may need to plot more than one graph for each $\omega$. Then, combining the solutions, $\phi$ ranges $(-\infty,+\infty)$ as in the classical case. Although there will be a singularity as $\phi$ goes through the critical value, the metric will be finite there, unlike at the classical counterparts. It is not sensible, in the context of semi-classical dilaton gravity, to talk about how objects could pass through the singularity. However, one can find solutions for values of $\phi$ above critical, and thus approach the singularity from `either side' as far as the dilaton is concerned. As one approaches the critical value however, the equations which are derived here from the action(\ref{eq:gen}) no longer represent the quantum theory of the action. This is because the graviton-dilaton loops become comparable to the large N matter field corrections. Thus, it is not clear how one should interpret the semi-classical singularity. One cannot make definite statements because we do not have a perturbative expansion or exact theory which indicates whether or not it persists. On either side of the singularity, however, the equations should be reliable. In \cite{BGHS}, purely super-critical solutions were considered interesting, and such a solution, that of constant curvature space, was presented. There exist four-dimensional extremal black hole solutions for which the asymptotic dilaton value is super-critical\cite{BAN}. For this reason, this author believes that it is useful to include the super-critical solutions here, even if some of the section on planar general relativity is superceded, because the quantum theory may turn out not to have a well-defined evolution through the critical value. The semi-classical appearance of a singularity is a limitation. It cannot be integrated through numerically and there is no contact between the sub- and super-critical dilaton regions of spacetime. A smoothed singularity should be passed through by test particles. In that case it seems plausible that one consider how to paste together semi-classical spacetimes which display singularities which do not appear at infinitely strong dilaton coupling. Naturally, an objection is that one has to reapply boundary conditions for the super-critical region, and so the solutions may have nothing to do with the subcritical ones. However, it is plausible to apply equivalent conditions. This is supported by the fact that the dilaton field is continuous across the critical line when one considers attaching sub- and super-critical solutions. As long as one has a classical spacetime picture, it would seem difficult to go further than this. However, it is precisely this type of operation that one has in mind for evaporating black holes which develop baby universes. The spacelike boundary is removed and replaced with a region to the future, disconnected from the external space. The quantum-corrected cases are commented upon individually in the following. \vbox{ \[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{q1rhfig.ps}\\ \end{array}\] FIG. 5. The conformal factor at the semi-classical level for subcritical initial value.} \subsection{Sub-Critical Solutions} Since the critical value is at $\phi_{cr}=-\frac 1 2\log \omega$, the lowest two cases, $\omega=0,-1$ do not have such a critical value. But since there is a change of initial gradient of $\phi$ when $\phi_{0}=\frac 1 2 \log 2$, we use this to divide the cases, whilst the other cases, $\omega=\frac 1 2, 1, \frac 3 2$ are genuinely subcritical in the FIG. 4-6. Generally speaking, the subcritical corrected cases resemble their classical partners at weak coupling, which is why sub-critical solutions have received more attention. The differences become clearer as the critical value is approached from below. \vspace{0.3in} \centerline{\bf A. \ $\omega=-1; \phi_0=0$ } \smallskip This case covers the qualitative behaviour for all $0>\omega>-\infty$. At strong coupling toward negative $s$, the curvature goes to minus infinity at $s=-\infty$. To the right, $\phi\to -\infty$ and $\rho\to\infty$ at finite s, whilst the curvature is always negative and finite. Thus we have a timelike {\it classical type} singularity to the left, and timelike infinity to the right, rather like the extremal Reissner-Nordstrom spacetime in four dimensions. This does not differ globally from the classical case. \vspace{0.3in} \centerline{\bf B. \ $\omega=0$; $\phi_0=0$} \smallskip The classical and quantum types of singularity coincide, because the place where the quantum singularity occurs in the case of $\omega=0$ is at infinity, as in the classical case. This is again anti-de Sitter space. The global structure is the same as for the classical case, though the dilaton goes to infinity toward the left in a logarithmic fashion in the corrected case. Lemos and S\` a showed that the classical Jackiw-Teitelboim theory contains a non-singular black hole, and this would appear to apply also to the quantum-corrected case. This black hole has zero radiation density as $\omega=0$, and zero temperature, as Cadoni and Mignemi showed\cite{CM}. \vbox{ \[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{q1rfig.ps}\\ \end{array}\] FIG. 6. The curvature at the semi-classical level for subcritical initial value.} \vspace{0.3in} \centerline{\bf C. \ $\omega=\frac 1 2, 1, \frac 3 2 $} \smallskip For $\omega=\frac 1 2$, the critical value coincides with the important value $\phi=\frac 1 2\log 2$, (see \ref{eq:derphi}). Each of these cases has a spacelike singular black hole to the left, as do higher values of $\omega$. The singularities occur at finite $\phi$, so the $\phi$(FIG. 6) and $\rho$(FIG. 7) graphs are truncated behind the horizon. In the asymptotic region to the right, the behaviour is as classically as expected at weak coupling for sub-critical plots, quantum corrections being small; the cases $\omega=\frac 1 2$ and $\omega=1$ are asymptotically anti-de Sitter and flat respectively, whereas for $\omega>1$ the curvature goes to minus infinity logarithmically, giving a timelike singularity, though this fact is not clear from FIG. 6, but can easily be confirmed by plotting to larger $s$. This zero coupling timelike singularity seems somewhat unphysical. The $\omega=1$ case agrees with the work of Birnir {\it et al}\cite{BGHS}. {}From the generalised asymptotic expansion given in\cite{SHE}, \begin{equation} \rho=\log{2b} - {K+L\log s \over s^{2b}} + \Big[{K+L\log s \over s^{2b}}\Big]^2 + ... \label{eq:rho} \end{equation} where $K$,$L$ and $b$ are arbitrary constants to be determined, and \begin{equation} \phi = \rho - b\log s - c. \label{eq:phi} \end{equation} Using(\ref{eq:mass}), in terms of the coordinate $2x=b\log s+2c$, where c is a constant, the ADM mass is \begin{equation} M = e^{2c}(K + {1\over{2b}}L(x+\alpha))_{x\to \infty}. \label{eq:infty} \end{equation} where $\alpha$ is a constant. The mass is therefore formally infinite, as was seen in\cite{BGHS}. It is because of the correspondence with classical theory that we regard the sub-critical case as physically interesting. But this has meant that the super-critical case has been left uninvestigated, and the semi-classical singularity mysterious. In the following `the other side' of the singularity, on which the dilaton is super-critical, and which was originally termed the `Liouville Region', is considered. \vbox{ \[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{q3pfig.ps}\\ \end{array}\] FIG. 7. The dilaton at the semi-classical level for super-critical initial value.} \subsection{Super-Critical Solutions} The super-critical initial value is taken to be $\phi_{0}=\log 2$ in the following five cases, which are given in FIG.s 7-9. For the cases $\omega=1,\frac 3 2$, there is another region between the critical value and $\phi_{0}=\frac 1 2\log2$ in which the dilaton remains confined. This additional super-critical pair of solutions is given in FIG.s 10-12 for completeness. \vbox{\[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{q3rhfig.ps}\\ \end{array}\] FIG. 8. The conformal factor at the semi-classical level, for supercritical initial value.} \vspace{0.3in} \centerline{\bf A. \ $\omega=-1,0$} \smallskip Now the strong coupling region is toward positive $s$, and $\phi$ decreases slowly at negative $s$. The space is again anti-de Sitter for $\omega=0$, while for $\omega=-1$ the curvature varies from zero at positive $s$, to $-4$ at negative $s$, where the spacetime approaches the anti-de Sitter space of the $\omega=0$ case. \vspace{0.3in} \centerline{\bf B. \ $\omega=\frac 1 2$} \smallskip This is the only case which has a spacelike singularity; the other cases are disconnected from their critical values in this region. In particular, the critical value is $\phi_{cr}=\frac 1 2 \log 2$. To the left $\phi$ diverges but the curvature is negative and slowly varying, while to the right, there is a singularity hidden by a horizon. We return to this case below. \vbox{\[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{q3rfig.ps}\\ \end{array}\] FIG. 9. The curvature at the semi-classical level, for supercritical initial value.} \vspace{0.3in} \centerline{\bf C. \ $\omega=1,\frac 3 2$; $\phi_0>\frac 1 2 \log 2$ } \smallskip These have finite curvature everywhere in this region. Toward negative $s$, at strong coupling, the space is asymptotically flat, while toward positive $s$, as $\phi\to\frac 1 2\log 2$, the curvature $R\to -4$. These results are in FIG.s 7-9. \vspace{0.3in} \centerline{\bf D. \ $\omega=1,\frac 3 2$; $\phi_{cr}<\phi_0<\frac 1 2 \log 2$ } \smallskip One can also consider the region immediately above the singularity as far as $\phi$ is concerned for these parameter values. The two examples are qualitatively similar. They have a timelike singularity at negative $s$ as the dilaton descends toward the critical value, and have approximately constant negative curvature at positive $s$, where $\phi\to\frac 1 2\log 2$ and $\rho$ diverges. These comments are given graphically in FIG.s 10-12. \vbox to \vsize{\[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{q2pfig.ps}\\ \end{array}\] FIG. 10. The dilaton for $\omega=1,\frac 3 2$ at the semi-classical level, for initial value which is above critical but below $\frac 1 2 \log 2$. \bigskip \[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{q2rhfig.ps}\\ \end{array}\] FIG. 11. The conformal factor for $\omega=1,\frac 3 2$ at the semi-classical level, for initial value which is above critical but below $\frac 1 2 \log 2$.\vfil} \vbox{\[ \begin{array}{c} \epsfxsize=0.6\displaywidth \epsfbox{q2rfig.ps}\\ \end{array}\] FIG. 12. The curvature for $\omega=1,\frac 3 2$ at the semi-classical level, for initial value which is above critical, but below $\frac 1 2 \log 2$.} \subsection{Planar General Relativity} The classical core of this case has been discussed in\cite{LEM}. The classical solution given earlier is actually a solution of general relativity. In this case, there are only two regions necessary to cover all values of $\phi$ along the real line, because the critical value coincides with the value $\phi=\frac 1 2\log 2$, which was discussed earlier. Generally, in such cases where there exists a finite region $\phi_{cr}<\phi<\frac 1 2 \log 2$, there is a solution in which the dilaton is confined between these outside values so that a third initial value will be necessary for the dilaton to range the full real line. Consider the results given in FIGS. 4 and 7. The curve $\omega=\frac 1 2$ in FIG. 7 increases monotonically from the asymptotic region at the right until it reaches $\phi_{cr}= \frac 1 2\log 2$ when there is a singularity and the integration breaks down. In FIG. 4, the curve descends monotonically to this value and has the same gradient there. If one were to attach these two curves, the dilaton would be a continous monotonic function {\it through} the singularity. There would be a small discontinuity in the conformal factor there. For other, asymetric initial dilaton values, the dilaton curve is no longer smooth, but it remains monotonic and piecewise continuous. The reader is reminded, however, that the equations are not valid at the singularity, so this point may not be important in a fuller theory. The Penrose diagrams for the extended spacetime corresponds to the two sets of solutions are given in FIG. 13. Dashed regions are copies of their undashed counterparts. \setlength{\unitlength}{0.0050in}% \hspace{60mm} \begin{picture}(160,520)(320,260) \put(680,440){\line( 0, 1){160}} \put(520,440){\line( 0, 1){160}} \put(680,440){\line(-1, 1){160}} \put(520,440){\line( 1, 1){160}} \put(520,600){\line( 1, 0){160}} \put(523,597){\line( 1, 0){154}} \put(520,440){\line( 1, 0){160}} \put(523,443){\line( 1, 0){154}} \put(650,520){\makebox(0,0)[lb]{\raisebox{0pt}[0pt][0pt]{I}}} \put(595,565){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{II}}} \put(595,465){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{II'}}} \put(530,520){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{ I'}}} \put(530,400){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{$\phi=\frac 1 2 \log 2$}}} \put(685,500){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{ $\phi=+\infty$}}} \put(540,350){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{ Case 2.}}} \put(380,440){\line( 0, 1){160}} \put(220,440){\line( 0, 1){160}} \put(380,440){\line(-1, 1){160}} \put(220,440){\line( 1, 1){160}} \put(220,600){\line( 1, 0){160}} \put(223,597){\line( 1, 0){154}} \put(220,440){\line( 1, 0){160}} \put(223,443){\line( 1, 0){154}} \put(350,520){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{ I}}} \put(295,565){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{II}}} \put(295,465){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{II'}}} \put(230,520){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{ I'}}} \put(230,620){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{$\phi=\frac 1 2 \log 2$}}} \put(100,500){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{$\phi=-\infty$}}} \put(240,350){\makebox(0,0)[lb] {\raisebox{0pt}[0pt][0pt]{ Case 1.}}} \end{picture} \centerline{FIG. 13: Penrose diagrams of quantum-corrected planar general relativity.} \vspace{0.2in} At the spacelike singularities of both boxes, $\phi=\frac 1 2\log 2$. One could draw a line from the timelike infinity in case 1., region I. where $\phi=-\infty$, through the singularity, where one identifies with a point on the lower singularity of case 2., and on through region II', out to timelike infinity in region I' where $\phi\to\infty$. A plausible pasting together of the two spacetimes would be to place the box corresponding to case 2. on top of that of case 1. and identify the singularities where the dilaton is $\frac 1 2\log 2$. This is very literally `toy modelling'. The interpretation of the resulting single diagram is open to debate. As expected, there is now a singularity at finite coupling in the {\it middle} of the diagram. Semi-classically, the singularity is final. However, in quantum gravity, the singularity may be smoothed, and test particles may be able to pass through the region of high curvature. The diagram here suggests a strong curvature wormhole shrouded on either `side' by an horizon, at which the curvature goes through zero, and is asymptotically anti-de Sitter. An observer who begins in the asymptotic region I of case I could avoid the wormhole by constantly accelerating immediately to timelike infinity, when $\phi=-\infty$, staying in region I. Alternatively, he might remain stationary, in which case he would pass through the wormhole at $\phi=\frac 1 2 \log 2$, after which he could constantly accelerate so that he reached another timelike infinity in the second box, where $\phi=\infty$. \section{Conclusions} A general, two-dimensional model has been considered and solved numerically for static, equilibrium solutions. There are many configurations which depend on both the value of the parameter $\omega$ and on the initial value of the dilaton field at the origin. The classical solutions found, bore out the results of \cite{LEMSA}, and the semi-classical $\omega=1,\phi<\phi_{cr}$ case, those of \cite{BGHS}. The extreme Reissner-Nordstrom type solution $\omega=-1$, the Schwarzschild type black hole $\omega=1$, of low-energy string theory, the black hole with timelike anti-de Sitter infinity $\omega=\frac 1 2$, of planar general relativity, the spacetime with both timelike and spacelike singularities, $\omega=\frac 3 2$, and the non-singular Jackiw-Teitelboim black hole $\omega=0$, were all seen both classically and at the semi-classical level, where they in general represented black holes in the Hartle-Hawking equilibrium state. These sub-critical solutions are the most clearly physically interesting solutions since they correspond with their classical counterparts and four-dimensional analogues. Of all these solutions, the unique parameter value which yields a solution which has everywhere positive curvature and is asymptotically flat, is that of string theory, $\omega=1$. Super-critical solutions for all cases were also found, and for ($\omega=\frac 1 2$), it appeared plausible to paste the super-critical to the sub-critical solution. Then, as in classical theories, the dilaton ranges the full real line and is continuous across the singularity. This construction is not possible in classical theory since in that case, the singularity always occurs at divergent dilaton field. At the semi-classical level, the two regions are still divided by a curvature singularity. This singularity was not expected originally\cite{CGHS}, and was met with puzzlement when discovered in subsequent work \cite{RSTA,BAN,BGHS}. In the {\sl RST} model, the singularity is taken seriously as a `central' boundary, analogous to the origin in Schwarzschild spacetime. However, it is known that there are energy conservational problems at the endpoint\cite{PS}, which may be related to this potential misinterpretation. The equations which generate the singularity become inappropriate in its vicinity, but one can still consider sub-critical and super-critical solutions independently. The singularity is a modification to classical theory which may or may not go away in quantum gravity, or is generically spurious. Birnir {\it et al}\cite{BGHS} discussed the possibility of sailing through this mild singularity. Horowitz and Marolf \cite{HM} have recently discussed the behaviour of quantum test particles which have well-defined motion even in singular spacetimes. \newpage \section{Acknowledgements} I thank Stephen Hawking and Gary Gibbons for reading the paper, and for conversations and suggestions.
\section{Introduction} The energy range of present and future particle colliders opens a kinematic regime, where the observation of jet events containing large rapidity gaps can provide interesting physical insight into the underlying exchange mechanisms. The presence of rapidity gaps can also serve as triggering signal in high-mass scale physics \cite{BJOne,BJSurvivalProbability,DokshitzerKhozeEtAl}. Events containing large rapidity gaps have recently been observed at hadron-hadron collider \cite{FermilabGaps} and lepton-hadron collider \cite{HeraDISGaps,HeraPhotoproductionGaps}. However, the possibility of observing jets separated by large gaps is not limited to hadron-hadron or lepton-hadron colliders. In fact, lepton-lepton collisions can also generate these events \cite{EEGaps,LEPIIGaps}. Purely leptonic collisions provide an environment free of spectator interactions, and can give interesting complementary information to the rapidity gap physics of hadronic collisions. Photon initiated collisions should provide yet another environment for the observation of jet events with rapidity gaps. The initial photons could be either real (for instance, by using a laser-backscattering beams \cite{GinzburgKotkinSerboTelnov}), or quasi-real, as those produced in the photoproduction events in $ep$ and $e^+e^-$colliders via the Weis\"acker-Williams equivalent photon mechanism \cite{WeisackerWilliams}. In fact, in the DESY HERA case, the gap events occur both at the deep-inelastic regime \cite{HeraDISGaps} as well as the photoproduction regime \cite{HeraPhotoproductionGaps}. (See also the theoretical models in Ref. \cite{HeraGapModels}.) However, at the moment the experimental situation has focused on the low $|t|$ region, where the proton or its low-mass excited state (like a $\Delta$) propagates in the forward direction, escaping detection. In this paper we explore into the higher $|t|$ region, where the proton is broken and a hard jet with $p_T^2 \sim |t| \gg 1$ GeV$^2$ is generated from the broken proton. This corresponds to the double diffractive dissociation of the $\gamma p$ system. {}From the study of resolved photon processes \cite{Witten,LlewellynSmith,BrodskyEtAl} (for a review, see Ref. \cite{DreesGodbolePramana}), we know that for the generation of jet events with fragments in forward and backward beam direction, the initial lepton can themselves be treated as containing hadronic components. Thus, one can talk about quark and gluon contents of electrons. For these jet events, the initial leptons would behave like supplier of the partons for the hard subprocess, much like the initial hadrons in hadron-hadron colliders. It is well known \cite{DreesGodboleLinearColliders} that as the collision energy $\sqrt{s}$ increases, the cross section for the annihilation events at lepton colliders decreases like $1/s$ or at best $\ln s/s$. At the same time, the cross section for the simplest hard two-photon process, $e^+e^- \to e^+e^- q \bar q$ increases like $\ln^3 s$, for fixed transverse momentum of the quarks or fixed invariant mass of the $q\bar q$ pair. In fact, at high enough energies $\sqrt{s} \sim 1 \ {\rm TeV}$, the combination of the increased cross section for resolved photon processes and the enhanced photon flux due to beamstrahlung can lead to severe hadronic backgrounds at $e^+e^-$ supercolliders. The photon content of electron is suppressed by $\alpha_{\rm em}$, but enhanced by a logarithmic factor $\sim \ln(s/4m_{\rm e}^2)$. In the kinematic regime of our interest ($ \sqrt{s} > 300$ \ GeV), this means that an electron beam can be qualitatively visualized as carrying a photon beam with a flux $5 \sim 10$\% that of the electron. Now, about a fraction $\alpha_{\rm em} \sim 10^{-2}$ of the time, we find a parton (quark or gluon) inside a photon. That is, taking all effects, electrons carry hadronic partons at the $10^{-3}$ level. If we consider double resolved processes, this means a factor of $10^{-3} \times 10^{-3} = 10^{-6}$ reduction in the flux. However, the parton diffractive scattering cross section is rather large, and it is essentially controlled by the cut-off in transverse momentum. This should be contrasted to the $e^+e^-$ annihilation cross section, which is controlled by the total center-of-mass energy $\sqrt{s}$. While the annihilation cross section becomes very small at large $\sqrt{s}$, the diffractive cross section at fix transverse momentum cut increases with $\sqrt{s}$. Therefore, we would expect the resolved photon diffractive events to be produced at an reasonable level at future $e^+e^-$ colliders, despite the flux reduction. For the real $\gamma\gamma$ colliders, we expect cross sections much larger than the $e^+e^-$ case, because the photons now are not coming from the electrons, hence there is less suppression of parton flux. ( At high energies, such as is planned for the Next Linear Collider \cite{Waikoloa}, potentially one should also consider $W$ and $Z$ bosons as partons inside electrons. But we will limit our scope here to the resolved photon contribution. ) For the resolved photon mechanism of generating jet events containing large rapidity gaps, we expect the spectator partons inside the photon to interact with the spectator partons inside the opposite beam particle. That is, the situation is similar to the case of hadron-hadron collider. The soft spectator interaction can generate particles and spoil the rapidity gap. The rapidity survival probability $\langle S^2 \rangle$ is defined as the probability that no other interaction occurs beside the hard interaction of interest \cite{BJSurvivalProbability}. This probability is most readily estimated as an average over the hadron-hadron impact parameter $B$ \cite{BJSurvivalProbability}: \begin{equation} \langle S^2 \rangle = \frac{\int d^2 B f(B) S^2(B)} {\int d^2 B f(B)}, \end{equation} where $S^2(B)$ is the probability that the colliding hadrons do not interact inelastically, and $f(B)$ is the cross section for the hard collision of interest. Different estimates for $\langle S^2 \rangle$ in hadron collisions, based on a variety of phenomenological models, are presented in Ref. \cite{GotsmanLevinMaor}, where $\langle S^2 \rangle$ is estimated to be between $0.05$ and $0.2$. $\langle S^2 \rangle$ is expected to depend on the colliding energy, but only weakly on the size of the rapidity gap. In Ref. \cite{Fletcher} the author uses a Good-Walker model for diffraction, and obtains a much higher value for the survival probability ($44\%$ at Tevatron energies, and $33\%$ at $40$ TeV.) On a related issue, the authors in Ref. \cite{ButterworthEtAl} have used the HERWIG Monte Carlo program and found that in $\gamma\gamma$ and $\gamma p$ collisions, the mean number of hard interactions per event ranges from $1.04$ to $1.123$ for various particle colliders. To obtain the cross section of jet events with rapidity gaps, we must multiply the hard collision cross section by the survival probability, that is, \begin{equation} \sigma_{\rm gap} = \langle S^2 \rangle \sigma_{\rm hard}. \end{equation} Qualitatively, we expect the survival probability involving photon initial states to be approximately the same of the survival probability involving only hadronic initial states. This can be argued based on the vector meson dominance picture. However, it would be very interesting to study the difference of survival probabilities from hadronic and photonic initial states. The measurement of resolved photon gap events can clarify this difference. In this paper we will present only the hard cross sections, without taking into the soft physics of the survival probability. But it will be implicitly understood that in order to obtain the final cross sections, the factor $\langle S^2 \rangle$ must be multiplied. This paper is structured as follows. In Section II we will study the gap event cross section at HERA in the photon-proton double diffractive dissociation region. In Section III we analyze the situation at future linear collider energies ($0.5$ to $1.5$ TeV), and in Section IV we analyze the situation for $\gamma\gamma$ colliders for a similar region of energies. Finally, in Section V we give the conclusions. \section{Resolved Photon Gap Events at HERA} The mechanism of generating resolved photon gap events at HERA $ep$ collider is illustrated in Fig. 1. The partons that participate in the hard collision subprocess can be either quarks or gluons. We will consider the hard collision regime $|t| \gg 1$ GeV$^2$ where the perturbative pomeron is applicable. Notice that this kinematic regime differs from the previous gap event regimes in Ref. \cite{HeraDISGaps,HeraPhotoproductionGaps}, where the proton remains unbroken or is excited to a low-mass state, and propagates down the beam pipe, escaping detection. That is, the rapidity gaps observed so far in HERA are between the real or virtual photon fragments and the unbroken proton or its excited state. This situation is different from the gap events observed at hadron-hadron colliders \cite{FermilabGaps}, where the rapidity gap is observed between two measured jets. Here in our paper, we consider a situation similar to the hadron-hadron collision case. That is, hard jets are generated both from the photon fragmentation and the proton fragmentation regions. This corresponds to the double diffractive dissociation of the $\gamma p$ system, and the rapidity gap exists between the two observed hard jets. \begin{figure}[htbp] \begin{center} \leavevmode { \epsfxsize=3.00in \epsfbox{fig1.eps} } \end{center} \caption[*]{ Resolved photon mechanism for producing jet events with a large rapidity gap in $ep$ collision. The partons inside the photon and the proton undergo a hard scattering via the exchange of a perturbative QCD pomeron. } \label{Fig1} \end{figure} The diffractive scattering cross section for the quark-quark $t$-channel color-singlet exchange case has been obtained by Mueller and Tang \cite{MuellerTang} (see also Ref. \cite{MorePomeron}) by using the Balitsky-Fadin-Kuraev-Lipatov (BFKL) model \cite{BFKLPomeron}: \begin{equation} \frac{d\sigma_{qq}}{dt} = \bigl(\alpha_{\rm s} C_{\rm F} \bigr)^4 \frac{\pi^3}{4t^2} \frac{e^{2(\alpha_{\rm P}-1)y}} { \Bigl[ \frac{7}{2} \alpha_{\rm s} C_{\rm A} \zeta (3) y \Bigr]^3 }, \end{equation} where $\alpha_{\rm s}=\alpha_{\rm s}(-t)$ is the strong coupling constant, $y= \ln(-\hat{s}/t)$ is the rapidity interval between the two out-going partons as measured from the beam axis, where $\hat{s}$ is the total center-of-mass energy squared of the $qq$ system; $\alpha_{\rm P} = 1+(4\alpha_{\rm s}C_{\rm A}/\pi) \ln 2$ the slope of the pomeron trajectory, $\zeta(3)=1.20206\dots$, and $C_{\rm A} = 3$ and $C_{\rm F} = 4/3$ the values of the Casimir operators in the adjoint and fundamental representations of the $SU(3)$ group. For the case of gluon-gluon elastic scattering with color-singlet $t$-channel exchange, we need only to replace the $C_{\rm F}$ factor in $d\sigma_{qq}/dt$ by $C_{\rm A}$. That is, \begin{equation} \frac{d\sigma_{gg}}{dt} = \alpha_{\rm s} C_{\rm A} \frac{\pi^3}{4t^2} \frac{e^{2(\alpha_{\rm P}-1)y}} { \Bigl[ \frac{7}{2} \zeta (3) y \Bigr]^3 }. \end{equation} To obtain the total cross section, we must integrate the cross section of the hard collision weighed by the respective structure functions. \begin{eqnarray} \sigma_{ep} (s,m_{\rm cut}^2,Y_{\rm cut}) &=& \int_{(m_{\rm cut}^2, Y_{\rm cut})} dz dx_1 \ dx_2 \ dt \ f_{\gamma|e} (z,s) \cr & & P_{\gamma}(x_1,-t) \ P_p(x_2,-t) \ \frac{d\sigma_{gg}}{dt} \bigl( \hat{s} = z x_1 x_2 s, t \bigr). \label{EqHeraGapCrossSection} \end{eqnarray} In the above formula, $z$ is the momentum fraction of the incoming electron carried by the photon, $x_1$ and $x_2$ are the momentum fraction of the partons carried by the photon and proton, respectively. $\sqrt{s}$ is the total center-of-mass energy of the $ep$ system, $m_{\rm cut}^2$ is the minimum transverse momentum squared of the hard tagging jets, and $Y_{\rm cut}$ is the minimum rapidity interval separating the two hard jets. The number of photons carrying a fraction $z$ of the energy of an emitting electron in leading log approximation is given by $f_{\gamma|{\rm e}}(s,z)= \alpha_{\rm em}/(2\pi z) [1+(1-z)^2] \ln(s/4m_{\rm e}^2) $, with $\sqrt{s}$ the total center-of-mass energy of the colliding $ep$ system and $m_{\rm e}$ the electron mass \cite{WeisackerWilliams,BawaStirling}. However, this formula overestimates the direct $\gamma\gamma \to q \bar{q}$ contribution \cite{BhattacharyaSmithGrammer}. An improved expression including non-leading terms is \cite{BrodskyKonishitaTerazawa}: \begin{eqnarray} f_{\gamma|{\rm e}}(z,s)= \frac{\alpha_{\rm em}}{2 \pi z} &\Biggl\{& \Bigl[ 1 + (1-z)^2 \Bigr] \Bigl( \ln(s/4m_{\rm e}^2) -1 \Bigr) \cr &+& z^2 \Biggl[ \ln \frac{2 (1-z)}{z} + 1 \Biggr] + (2-z)^2 \ln \frac{2 (1-z)}{2-z} \Biggr\}, \end{eqnarray} This formula has been shown \cite{BhattacharyaSmithGrammer} to give accurate results not only for total (direct) jet rates but also for distributions. In Eq. (\ref{EqHeraGapCrossSection}), the quantities $P_{\gamma}(x_1,-t)$ and $P_p(x_2,-t)$ correspond to the parton structure functions of the photon and proton, respectively, and they are defined as \begin{eqnarray} P_{\gamma}(x,Q^2) &=& f_{g|\gamma} (x,Q^2) + \Biggl( \frac{C_{\rm F}}{C_{\rm A}} \Biggr)^2 \sum_q \Bigl[ \ f_{q|\gamma} (x,Q^2) + f_{\bar{q}|\gamma} (x,Q^2) \ \Bigr], \cr P_p(x,Q^2) &=& f_{g|p} (x,Q^2) + \Biggl( \frac{C_{\rm F}}{C_{\rm A}} \Biggr)^2 \sum_q \Bigl[ \ f_{q|p} (x,Q^2) + f_{\bar{q}|p} (x,Q^2) \ \Bigr]. \end{eqnarray} For the photon structure functions, we have the identity $f_{q|\gamma} (x,Q^2)= f_{\bar{q}|\gamma} (x,Q^2)$. There exist many parametrizations for the parton distributions inside proton and photon. For the photon distribution functions, we will limit ourselves to the Drees-Godbole (DG) parametrization \cite{DreesGodboleParametrization} and to the Gl\"uck-Reya-Vogt (GRV-LO) parametrization \cite{GluckReyaVogtParametrization}. (We thank the authors for providing the programs.) Numerically, the largest uncertainty in our calculation comes from the gluon density inside photon, where little experimental result is available. The quark densities are better understood. See the recent results by the TRISTAN's collaborations TOPAZ \cite{TOPAZ} and AMY \cite{AMY} on the measurement of the photon structure function $F_2^\gamma(x,Q^2)$. It is worth mentioning that the gluon density of the photon in the GRV-LO parametrization is consistent with the recent measurement by the H1 Collaboration at HERA \cite{H1PhotonStructure}. For the proton structure function, we choose the CTEQ 2'M parametrization from the CERN PDFLIB routine library \cite{PlothowBesch}. We use $Q^2=-t$ for the scale of the photon and proton structure functions. We perform the numerical integration with the Monte Carlo integration program VEGAS \cite{Vegas}. The integration limits are \begin{eqnarray} && 0 \le z, x_1, x_2 \le 1, \cr && m_{\rm cut}^2 \le |t| \le x_1 x_2 e^{-Y_{\rm cut}} s, \end{eqnarray} where we integrate over all events with jet transverse momentum larger than $m_{\rm cut}^2$ and a rapidity separation between the jet centers larger than $Y_{\rm cut}$. It should be pointed out that due to the hadronization effect, the hadron fragments typically scatter around the jet centers, within a circle of radius $\sim 0.7$ units of rapidity \cite{BJSurvivalProbability}. Hence the observed effective gap is $Y_{\rm eff} \sim Y - 1.4$. \begin{figure}[htbp] \begin{center} \leavevmode { \epsfysize=5.00in \epsfbox{fig2.eps} } \end{center} \caption[*]{ Resolved photon gap event cross section at HERA, for various values of the transverse moment cut $m_{\rm cut}$. The solid lines are obtained with the Gl\"uck-Reya-Vogt parametrization of the photon structure functions, and the broken lines with the Drees-Godbole parametrization. } \label{Fig2} \end{figure} In Fig. 2 we present the result of the calculation. For $Y_{\rm cut}=4$ and $m_{\rm cut}=5$ \ GeV, even taken into account the survival probability consideration and uncertainty for the gluon distribution inside the photon, the gap events should still be produced at an observable rate. As we stated before, gap events have been observed at HERA both in the deep-inelastic regime \cite{HeraDISGaps} and in the photoproduction regime \cite{HeraPhotoproductionGaps}, where the rapidity gap exists between the photon (real or virtual) fragmentation region and the forward, undetected proton (or its excited state.) It would be interesting to verify the existence of rapidity gap between two observed hard jets at HERA, and study the dependence of these gap event cross sections on the rapidity interval $Y_{\rm cut}$ and on the transverse moment cut $m_{\rm cut}^2$, and compare with our calculation here. We should note, however, that smaller rapidity gaps can also arise from random fluctuation of the fragments of hadronization process. A systematic study of rapidity gap physics here and in other environments \cite{FermilabGaps} will gradually allow better understanding of the relative importance of gap events from perturbative mechanisms and from random fluctuations, as well as insight to the survival probability involving photons. It is important to point out, though, that our current knowledge of the photon structure functions is rather imprecise. This is especially true for the gluon content of photon, which numerically forms the dominant contribution to the gap event cross section. Therefore, an uncertainty of half an order of magnitude above or below the calculated curves would not be unreasonable. Hopefully the gluon content of the photon can be better measured at HERA in the future \cite{H1PhotonStructure}. \section{Resolved Photon Gap Events at NLC} \begin{figure}[htbp] \begin{center} \leavevmode { \epsfxsize=3.00in \epsfbox{fig3.eps} } \end{center} \caption[*]{ Resolved photon mechanism for producing jet events with a large rapidity gap in $e^+e^-$ collision. The partons inside the photons undergo a hard scattering via the exchange of a perturbative QCD pomeron. } \label{Fig3} \end{figure} The production mechanism is illustrated in Fig. 3. Currently, the energy projected for future $e^+e^-$ colliders is in the range $0.5$ to $1.5$ TeV. At this energy range, the resolved photon contributions should be substantial. However, due to beamstrahlung background, the detectors for these colliders are not expected to be sensitive to jet production in the forward or backward direction. That is, many of the gap events will not be detected. We will analyze the production cross section of gap event by taking into account also the detector limitation. Here we will assume that the detector is only capable of observing hard jets produced in the rapidity region $[-\eta_{\rm det}, \eta_{\rm det}]$. Given this limitation, the observable rapidity gap events can be classified into three cases: (a) both hard jets are observed, (b) only the forward hard jet is observed, and (c) only the backward hard jet is observed. These situations are illustrated in Fig. 4. In the cases (b) and (c), since one of the hard jets is not measured, it is not possible to know the true size of the rapidity gap. For these cases, we define the empirical rapidity gap as the size of the rapidity interval between the hard jet and the detector limit on the opposite side. Mathematically, if $y_1$ and $y_2$ represent the rapidities of the forward and the backward hard jet, then we define the empirical gap to be \begin{equation} Y = {\rm Min} \left\{ y_1 - y_2, y_1 + \eta_{\rm det}, \eta_{\rm det} - y_2 \right\} , \end{equation} Naturally, the size of the rapidity gap cannot exceed the detector range: $Y < 2 \eta_{\rm det}$. \begin{figure}[htbp] \begin{center} \leavevmode { \epsfxsize=4.00in \epsfbox{fig4.eps} } \end{center} \caption[*]{ Three possible cases for the observation of jet events containing rapidity gaps: (a) both hard jets are observed, (b) only the forward jet is observed, and (c) only the backward jet is observed. In the cases of (b) and (c), the gap size is measured from the jet center to the rapidity edge of the detector on the opposite side. } \label{Fig4} \end{figure} To integrate the event cross section, we take into account the kinematic cuts from 1) detector limitation: $\eta_{\rm det}$, 2) rapidity gap cut: $Y_{\rm cut}$, and 3) transverse momentum cut: $m_{\rm cut}^2$. \begin{eqnarray} \sigma_{e^+e^-} (s,m_{\rm cut}^2,Y_{\rm cut},\eta_{\rm det}) &=& \int_{(m_{\rm cut}^2, Y_{\rm cut}, \eta_{\rm det})} dz_1 dz_2 dx_1 \ dx_2 \ dt \ f_{\gamma|e} (z_1,s) f_{\gamma|e} (z_2,s) \cr & & P_{\gamma}(x_1,-t) \ P_{\gamma}(x_2,-t) \ \frac{d\sigma_{gg}}{dt} \bigl( \hat{s} = z_1 z_2 x_1 x_2 s, t \bigr). \label{EqNLCGapCrossSection} \end{eqnarray} The photon momentum fractions inside the electron and the positron are respectively $z_1$ and $z_2$. Other quantities that appear in this formula have been explained in the previous section. The integration momentum fractions are constrained to: $0 \le z_1, z_2, x_1, x_2 \le 1.$ The gap and transverse momentum constraints impose the following limits for the $t$ integration. For the case (a), \begin{equation} {\rm Max} \left\{ \begin{array}{c} m_{\rm cut}^2 \\ x_1^2 z_1 z_2 e^{-2\eta_{\rm det}} s \\ x_2^2 z_1 z_2 e^{-2\eta_{\rm det}} s \\ \end{array} \right\} \ < \ \left| t \right| \ < \ x_1 x_2 z_1 z_2 e^{-Y_{\rm cut}} s . \label{EqLimitA} \end{equation} For the case (b), \begin{equation} {\rm Max} \left\{ \begin{array}{c} m_{\rm cut}^2 \\ x_1^2 z_1 z_2 e^{-2\eta_{\rm det}} s \\ \end{array} \right\} \ < \ \left| t \right| \ < \ {\rm Min} \left\{ \begin{array}{c} x_1^2 z_1 z_2 e^{-2(Y_{\rm cut}-\eta_{\rm det})} s \\ x_2^2 z_1 z_2 e^{-\eta_{\rm det}} s \end{array} \right\} . \label{EqLimitB} \end{equation} And for the case (c), \begin{equation} {\rm Max} \left\{ \begin{array}{c} m_{\rm cut}^2 \\ x_2^2 z_1 z_2 e^{-2\eta_{\rm det}} s \\ \end{array} \right\} \ < \ \left| t \right| \ < \ {\rm Min} \left\{ \begin{array}{c} x_1^2 z_1 z_2 e^{-\eta_{\rm det}} s \\ x_2^2 z_1 z_2 e^{-2(Y_{\rm cut}-\eta_{\rm det})} s \\ \end{array} \right\} . \label{EqLimitC} \end{equation} We include all three contributions in our integration of the event cross sections. In Fig. 5 and Fig. 6 are the results of the gap event cross sections, for $0.5$ and $1.5$ TeV center-of-mass energy. We consider three values for the maximum detector rapidity: $\eta_{\rm det} = 2, 3, 4$, which means that the maximum detectable rapidity gaps are respectively $Y=4, 6, 8$. The curves are plotted for two values of the transverse momentum cut: $m_{\rm cut} = 5, 10$ GeV. \begin{figure}[htbp] \begin{center} \leavevmode { \epsfysize=5.00in \epsfbox{fig5.eps} } \end{center} \caption[*]{ Gap event cross section for $e^+e^-$ collider at $E_{\rm cm}=0.5$ TeV, for various values of transverse momentum cut $m_{\rm cut}=5,10$ GeV and detector rapidity limits $\eta_{\rm det}=2,3,4$. The maximum observable rapidity gaps are $2 \eta_{\rm det} = 4,6,8$ and are indicated by the dot-dashed vertical lines. The solid lines are obtained by using the Gl\"uck-Reya-Vogt parametrization of the photon structure functions, and the broken lines by using the Dree-Godbole parametrization. } \label{Fig5} \end{figure} \begin{figure}[htbp] \begin{center} \leavevmode { \epsfysize=5.00in \epsfbox{fig6.eps} } \end{center} \caption[*]{ Same as Fig. 5, but at a center-of-mass energy of $E_{\rm cm} = 1.5$ TeV. } \label{Fig6} \end{figure} With the projected luminosity for the future linear colliders $\sim 10^{34}$ [cm$^{-2}$s$^{-1}$] , the detectors should be sensitive to the physics at cross section level of $\sim 1$ [fb]. Hence, unless the detector has very narrow range of rapidity, gap events with $Y_{\rm cut} = 4$ and $m_{\rm cut} = 5$ GeV should be produced at an observable rate, this even when the survival probability is taken into account. Notice that in going from $0.5$ to $1.5$ TeV, the gap event cross section increases by about an order of magnitude. (Naturally, we also have to keep in mind the uncertainty from the photon structure functions.) In Ref. \cite{LEPIIGaps}, various mechanisms for the production of rapidity gap events at LEP-II have been analyzed. These mechanisms can be characterized as the annihilation of $e^+e^-$ into two gauge bosons, which subsequently decay into jet pairs. As opposed to the resolved photon mechanism studied here, in the annihilation mechanisms all the beam energy goes into the production of the hadronic jets. In principle, it is possible to distinguish these two mechanisms, by measuring the presence or absence of the $e^+e^-$ in the forward or backward direction. (Calorimetry may also help, although the lepton colliders are not expected to be sensitive to forward and backward jets due to the background problems). In practice, this distinction may not always be feasible. In terms of orders of magnitude, the annihilation mechanisms like $e^+e^- \to \gamma^*\gamma^* \to jets$ and $e^+e^- \to \gamma^* Z \to jets$ may be produced at a competing level with the resolved photon cases (at least for the $0.5$ TeV machine). Also, there are other mechanisms of producing gap events, such as coming from $W$ and $Z$ bosons, via annihilation or resolved mechanisms. In summary, there is a rich phenomenology still to be studied. We limit our scope here only to the resolved photon contribution, and postpone a more comprehensive analysis of rapidity gap jet events at NLC for the future. \section{Resolved Photon Gap Events at $\gamma\gamma$ Collider} As opposed to the $e^+e^-$ case, there is no photon flux suppression for real photon collisions. Therefore, we expect the resolved photon events to provide a much larger cross section for gap events. \begin{figure}[htbp] \begin{center} \leavevmode { \epsfxsize=3.00in \epsfbox{fig7.eps} } \end{center} \caption[*]{ Resolved photon mechanism for producing jet events with a large rapidity gap in $\gamma\gamma$ collision. The partons inside the photons undergo a hard scattering via the exchange of a perturbative QCD pomeron. } \label{Fig7} \end{figure} The production mechanism is illustrated in Fig. 7. As in the case of $e^+e^-$ collider, we will assume a rapidity range of $[-\eta_{\rm det}, \eta_{\rm det}]$ for the detector. Although theoretically a $\gamma\gamma$ collider should have little beamstrahlung effects, hence the detectors should be able to observe jets near the forward and backward direction, in practice this may not be true. There remains serious technological challenge to the conversion of $e^+e^-$ colliders into $\gamma\gamma$ colliders. In particular, the distance between the laser-backscattering points and the $\gamma\gamma$ collision point may not be large enough for the deflection of the remnant $e^+e^-$ beams \cite{Heusch}. We also consider $0.5$ to $1.5$ TeV as the range for the center-of-mass energy. The event cross section is \begin{equation} \sigma_{\gamma\gamma} (s,m_{\rm cut}^2,Y_{\rm cut},\eta_{\rm det}) = \int_{(m_{\rm cut}^2, Y_{\rm cut}, \eta_{\rm det})} dx_1 \ dx_2 \ dt \ P_{\gamma}(x_1,-t) \ P_{\gamma}(x_2,-t) \ \frac{d\sigma_{gg}}{dt} \bigl( \hat{s} = x_1 x_2 s, t \bigr). \label{EqGammaGammaGapCrossSection} \end{equation} And the integration limits for the $t$ variable are similar to those ones given in Eq. (\ref{EqLimitA}), (\ref{EqLimitB}) and (\ref{EqLimitC}), upon substituting $z_1=z_2=1$. \begin{figure}[htbp] \begin{center} \leavevmode { \epsfysize=5.00in \epsfbox{fig8.eps} } \end{center} \caption[*]{ Gap event cross section for $\gamma\gamma$ collider at $E_{\rm cm}=0.5$ TeV, for various values of transverse momentum cut $m_{\rm cut}=5,10$ GeV and detector rapidity limits $\eta_{\rm det}=2,3,4$. The maximum observable rapidity gaps are $2 \eta_{\rm det} = 4,6,8$ and are indicated by the dot-dashed vertical lines. The solid lines are obtained by using the Gl\"uck-Reya-Vogt parametrization of the photon structure functions, and the broken lines by using the Dree-Godbole parametrization. } \label{Fig8} \end{figure} \begin{figure}[htbp] \begin{center} \leavevmode { \epsfysize=5.00in \epsfbox{fig9.eps} } \end{center} \caption[*]{ Same as Fig. 8, but at a center-of-mass energy of $E_{\rm cm} = 1.5$ TeV. } \label{Fig9} \end{figure} The results for the cross sections are plotted in Fig. 8 and Fig. 9. We can see that compared to the $e^+e^-$ case, the cross sections are now two to three orders of magnitude larger. Hence the gap events will be produced copiously at $\gamma\gamma$ colliders. This would provide the ideal environment of the study of survival probability involving photon initial states. \section{Conclusions} We have seen that resolved photons provides a mechanism of producing jet events containing large rapidity gaps, and we have analyzed the event cross section at HERA $ep$ collider and future $e^+e^-$ and $\gamma\gamma$ colliders. We have seen that in all three cases the event cross section are at the reach of the experiments. In the case of HERA, it would be interesting to observe the existence of rapidity gaps between two hard jet systems (one in the forward direction and the other one in the backward direction), and analyze the dependence of the cross section on the rapidity gap cut $Y_{\rm cut}$ and on the transverse momentum cut $m_{\rm cut}$. This would provide a first look into the survival probability involving photon-hadron collision. In the case of $\gamma\gamma$ collision we have seen that the event cross section becomes two to three orders of magnitude larger than the $e^+e^-$ case. We have also seen that the resolved photon gap events increase significantly with the total center-of-mass energy. The observation of these events will allow the study of the perturbative QCD pomeron physics in environments alternative to the hadron-hadron colliders, provide insight into the survival probability physics of photons, and also allow further understanding on the relative importance of gap events coming from random fluctuation of hadronization effects and from perturbative mechanisms. We very especially thank Wai-Keung Tang, for all the help received during the preparation of this work. We also thank Ina Sarcevic, Stanley Brodsky and Clemens A. Heusch for helpful conversations, and M. Drees, M. Gl\"uck, E. Reya and A. Vogt for providing the subroutines for the photon structure functions. This work was supported by U.S. Department of Energy Grants No. DE--FG03--93ER40792.
\section*{Acknowledgments} The authors thank the Graduiertenkolleg ``Mathematik im Bereich ihrer Wechselwirkung mit der Physik'' of the Ludwig Maximilian University Munich , in particular Prof. Batt and Prof. Schottenloher, for the possibility of a stay of T. A. at Munich where most of the ideas took form. G.H. is also indebted to Prof. Baum, Prof. Friedrich, Prof. Petry and W. Posch for helpful discussions on exotic spinors. T.A. thanks also Prof. Keiper for useful discussions.
\section*{Acknowledgements} We thank Y. Iwasaki, K. Kanaya and T. Yoshi\'e for showing us their unpublished data and for useful discussions. Numerical calculations for the present work have been carried out at Center for Computational Physics and on VPP500/30 at Science Information Processing Center, both at University of Tsukuba. This work is supported in part by the Grants-in-Aid of the Ministry of Education(Nos. 04NP0701, 06640372). \newpage
\section{Introduction} \indent Nonlinear differential equations with fixed critical points define a natural extension \cite{PaiOeuvres} of linear equations. Let us recall that a singular point is said {\it critical} if several determinations of the solution are permuted around it, and {\it movable} (contrary {\it fixed}) if its location depends on the initial conditions. The {\it Painlev\'e property} (PP) of a differential equation, is defined \cite{PaiOeuvres} as the absence of movable critical points in the general solution of the differential equation. The {\it Painlev\'e test} \cite{Chamonix1994} is the set of all methods to build necessary conditions for a differential equation to possess the PP, without guarantee on their sufficiency. The most powerful of them is the ``$\alpha-$method'' of Painlev\'e \cite{PaiBSMF}, but its difficulty is differential since at each step one must integrate a differential equation. All other methods (Kowalevskaya \cite{Kowalevskaya1889,ARS1980}, Bureau \cite{Bureau1939}, Conte, Fordy and Pickering \cite{FP1991,CFP1993}) have only an algebraic difficulty and therefore are easy to automatize. But all those algebraic methods share a restriction, which this Letter proposes to remove, thus adding to the Painlev\'e test a new algebraic method which very often allows to conclude to a failure more rapidly than with the old algebraic methods. In section \ref{sectionPerturbative}, we recall the lemma of Painlev\'e which establishes the perturbative framework, then we define our extension to the Painlev\'e test. Section \ref{sectionChazy} is a pedagogical example quite useful to explain the method. Two examples are then considered, one from mathematics section \ref{eqsectionBureau}, the second one from physics section \ref{sectionBianchi}. Both exhibit movable logarithms and provide a shorter proof of failure of the test than before. \section{The perturbative framework \label{sectionPerturbative}} \indent All the methods of the Painlev\'e test \cite{Chamonix1994}, whether differential (Painlev\'e) or algebraic (Kowalevskaya, Bureau, Conte {\it et al.}), are based on the following lemma of Painlev\'e \cite{PaiBSMF}. {\it Lemma}. Consider an algebraic differential equation of order $N$ \begin{equation} {\bf E}(x,{\bf u},\varepsilon)=0, \end{equation} (the boldface characters mean multicomponent) depending analytically on a small complex parameter $\varepsilon$, defined in the canonical form of Cauchy \begin{equation} {\hbox{d} {\bf u} \over \hbox{d} x}={\bf K}[x,{\bf u},\varepsilon],\ x \in {\cal C},\ {\bf u} \in {\cal C}^N,\ \varepsilon \in {\cal C}, \label{eqLemma} \end{equation} with ${\bf K}$ assumed holomorphic at $\varepsilon=0$. If the general solution of (\ref{eqLemma}) has no movable critical points for every $\varepsilon$ except maybe for $\varepsilon=0$, then \begin{itemize} \item{} the point $\varepsilon=0$ is no exception, i.e.~the general solution has also no movable critical points there, \item{} every coefficient of the Taylor series of ${\bf u}$ \begin{equation} {\bf u}=\sum_{n=0}^{+ \infty} \varepsilon^n {\bf u}^{(n)} \label{eqPerturbu} \end{equation} has no movable critical points. \end{itemize} A detailed proof of this lemma has been established by Bureau \cite{Bureau1939,BureauI}, in place of the too short proof by Painlev\'e, who merely considered it as an immediate consequence of the classical theorem of perturbations of Poincar\'e. Let us introduce the notation \begin{equation} {\bf E}(x,{\bf u},\varepsilon) \equiv\sum_{n=0}^{+ \infty} \varepsilon^n {\bf E}^{(n)}(x,{\bf u}^{(0)},\dots,{\bf u}^{(n)})=0 \end{equation} in which the equation ${\bf E}^{(0)}(x,{\bf u}^{(0)}) = 0$ is nonlinear and every equation \hfill \par \noindent ${\bf E}^{(n)}(x,{\bf u}^{(0)},\dots,{\bf u}^{(n)})=0,n\ge 1$ is linear inhomogeneous in ${\bf u}^{(n)}$. Consider an equation ${\bf E}(x,{\bf u})=0$ independent of $\varepsilon$, a case in which the lemma still applies (ref.~\cite{Goursat} vol.~III). The equations ${\bf E}(x,{\bf u})=0$ and ${\bf E}^{(0)}(x,{\bf u}^{(0)}) = 0$ are then the same, the equation ${\bf E}^{(1)}=0$ is the linearized of ${\bf E}$ at ${\bf u}^{(0)}$ and equations ${\bf E}^{(n)}=0, n \ge 2,$ only differ from ${\bf E}^{(1)}=0$ in the r.h.s., \begin{eqnarray} n=0:\ {\bf E}^{(0)} & \equiv & {\bf E}(x,{\bf u}^{(0)}) = 0, \label{eqNL0} \\ n=1:\ {\bf E}^{(1)} & \equiv & {\bf E}'(x,{\bf u}^{(0)}) {\bf u}^{(1)} = 0, \label{eqLin0} \\ \forall n \ge 2:\ {\bf E}^{(n)} & \equiv & {\bf E}'(x,{\bf u}^{(0)}) {\bf u}^{(n)} + {\bf R}^{(n)}(x,{\bf u}^{(0)},\dots,{\bf u}^{(n-1)}) = 0. \label{eqOrdern} \end{eqnarray} Suppose one knows a solution ${\bf u}^{(0)}$ which is global and without movable critical points, but which depends on an insufficient number $M<N$ of arbitrary parameters. We require, first, that its perturbation (\ref{eqPerturbu}) in the neighborhood of $\varepsilon=0$ represents the general solution, secondly, that each ${\bf u}^{(n)}$, in particular ${\bf u}^{(1)}$ in this Letter, be without movable critical points. Painlev\'e showed (ref.~\cite{PaiBSMF} p.~209 note 1) that $M$ of the $N$ solutions of the linearized equation (\ref{eqLin0}) are already known, these are the derivatives of ${\bf u}^{(0)}$ with respect to its $M$ parameters, evidently without movable critical points. Satisfying the first point is not so evident. Indeed, a peculiarity of nonlinear differential equations is that they sometimes possess, in addition to a general solution and its particular solutions, other solutions called singular solutions. Those are impossible to obtain from the general solution by giving special values to the arbitrary integration constants. For instance, the equation \cite{ChazyThese} \begin{equation} u'''=2 u' u'' + 2 i u'' \sqrt{u'' - u'^2 - 1} \end{equation} has for general solution: $u=e^{2 c_1 x + c_2} + (c_1^2 - 1) x /(4 c_1) + c_3$ and for singular solution: $u=- \mathop{\rm Log}\nolimits \cos (x + C_1) + C_2$, showing the absence of correlation \cite{ChazyThese} between the structure of singularities of the general solution and that of the possible singular solutions. Singular solutions are for sure excluded, as requested by the definition of the PP and the application of the above lemma, if the equation ${\bf E}^{(1)}=0$ is exactly of order $N$. Indeed, the singular solutions, which generalize the notion of envelope, satisfy a differential equation with an order smaller than $N$. But the mean used by the available algebraic methods to satisfy this requirement is to ask that the equation (\ref{eqLin0}) possesses exactly $N$ Fuchs indices [ for the basic vocabulary (singular regular, singular irregular, Fuchsian, non-Fuchsian, essential singularity, Fuchs indices) and methods concerning linear differential equations in the complex plane, the reader can refer to the classical book of Ince \cite{Ince}, chapters XV to XVII ] at the movable poles $x_0$ of ${\bf u}^{(0)}$ located at a finite distance, which implicitly discards possible non-Fuchsian solutions, even if they have no movable critical points. Let us briefly recall the differences between Fuchsian and non-Fuchsian for our purpose which is the question of local singlevaluedness. Near a regular singular point $x=0$, there exist $N$ linearly independent solutions \begin{equation} x^{\lambda_i} \sum_{j=0}^{m_i} \varphi_{ij}(x) (\mathop{\rm Log}\nolimits x)^{j},\ i=1,N \label{eqFundamentalSetFuchs} \end{equation} in which the $\lambda_i$'s are complex numbers (the Fuchs indices), $m_i$ positive integers (their multiplicity), $\varphi_{ij}$ converging Laurent series of $x$ with finite principal parts. The necessary and sufficient condition of local singlevaluedness of the general solution of the linear equation is: $\lambda_i$ all integer, no $\mathop{\rm Log}\nolimits$ terms. Near an irregular singular point $x=0$, there exist $N$ linearly independent solutions \begin{equation} e^{Q_i(1/z_i)} x^{s_i} \sum_{j=0}^{m_i} \varphi_{ij}(z_i) (\mathop{\rm Log}\nolimits x)^{j},\ z_i=x^{1/q_i},\ i=1,N \label{eqFundamentalSetNonFuchs} \end{equation} in which $q_i$ is a positive integer, $Q_i$ a polynomial, $s_i$ a complex number, $\varphi_{ij}$ a {\it formal} Laurent series with finite principal part. The question of local singlevaluedness of the general solution cannot be settled so easily, because formal series are generically divergent. In this Letter, we go back to the lemma of Painlev\'e and we implement two features. The {\it first feature} is to also accept that $x_0$ be non-Fuchsian for the linearized equation (\ref{eqLin0}): we require the existence of a fundamental set of solutions ${\bf u}^{(1)}$ of ${\bf E}^{(1)}=0$ which are locally single valued near $\chi=x-x_0=0$, and the same property for a particular solution ${\bf u}^{(n)}$ of each ${\bf E}^{(n)}=0,n \ge 2$. The {\it second feature} needs an additional assumption, namely that the given solution ${\bf u}^{(0)}$ be known globally (in closed form). The singular points of equation (\ref{eqLin0}) can be classified into three types: \begin{enumerate} \item the fixed singularities of equation (\ref{eqNL0}) located at a finite distance, \item the movable poles $x_0,x_1,\dots$ of ${\bf u}^{(0)}$ located at a finite distance, \item the point at infinity, which is fixed. \end{enumerate} All three types can be studied because ${\bf u}^{(0)}$ is known globally. The first type of singularities must not be studied because the PP allows any behaviour around them. Each point $x_0,x_1,\dots$ of the second type must be studied, with the same requirements than in the first feature. As to the third type, namely the point at infinity, it should generally not be tested for it is fixed. However, in case there are no critical singularities of the first type (fixed critical singularities at a finite distance), one can also require the existence of a fundamental set of solutions locally single valued near $\infty$; in particular, it may happen the favourable event that, while the point $x_0$ is non-Fuchsian, the point $\infty$ is Fuchsian and allows to conclude about {\it global} singlevaluedness. \medskip Our addition to the Painlev\'e test can be detailed as follows. \begin{enumerate} \item Assume a given solution ${\bf u}^{(0)}$ which is global and has at least one movable pole at a finite distance denoted $x_0$. \item Require the linearized equation at ${\bf u}^{(0)}$ to have exactly order $N$, so as to discard singular solutions. \item Near every movable pole $x_0$ of ${\bf u}^{(0)}$ located at a finite distance, build a fundamental set of solutions ${\bf u}^{(1)}$ and require it to be locally single valued. \item In case of at least one non-Fuchsian point among $x_0,x_1,\dots$ at the preceding step, and if there is no fixed critical singularity at a finite distance, build a fundamental set of solutions ${\bf u}^{(1)}$ near $\chi=\infty$ and require it to be locally single valued. \item At each higher perturbation order $n \ge 2$, build similarly particular solutions ${\bf u}^{(n)}$ and require the same properties. \end{enumerate} We will call a {\it family} {\it Fuchsian} or {\it non-Fuchsian} according to the nature (Fuchsian or non-Fuchsian) of the singular point $\chi=0$ of the linearized equation. The most likely event to occur in our extension of the test, leading to its failure, is the detection of a movable logarithm coming from two integer Fuchs indices at the point $\chi=\infty$, of course under the condition of absence of any fixed critical singularity at a finite distance. The information (pass or fail) which we propose to extract is \begin{enumerate} \item not accessible to the method of Kowalevskaya, which accepts the Fuchsian families as well as the non-Fuchsian ones but does not consider the perturbation (\ref{eqPerturbu}), \item not accessible to the method of Bureau, which rejects the non-Fuchsian families (because the representation by divergent series forbids to conclude) and does not consider the perturbation (\ref{eqPerturbu}), \item accessible to the method of Conte, Fordy and Pickering, which rejects the non-Fuchsian families for the same reason than Bureau and extracts the information from Fuchsian families only at the expense of a {\it Fuchsian} perturbative computation which may sometimes be long. \end{enumerate} \section{An explanatory example: Chazy's class III ($N=3,M=2$) \label{sectionChazy}} \indent The simplified (i.e.~scaled) equation of class III of Chazy \cite{ChazyThese} \begin{equation} E(x,u) \equiv u''' - 2 u u'' + 3 u'^2 = 0. \label{eqChazy} \end{equation} admits the global two-parameter solution \begin{eqnarray} u^{(0)} & = & c \chi^{-2} - 6 \chi^{-1},\ \chi=x-x_0,\ (x_0,c) \hbox{ arbitrary}. \label{eqChazyOrder0} \end{eqnarray} For that equation, this solution arises from the {\it local} search for all the families of movable singularities $ u \sim u_0 \chi^p,\ E \sim E_0 \chi^q,\ \chi=x-x_0 \to 0$ represented by Laurent series with a finite principal part \begin{eqnarray} p=-1,\ q=-4,\ u^{(0)} & = & -6 \chi^{-1}, \label{eqChazyFamily1} \\ p=-2,\ q=-6,\ u^{(0)} & = & c \chi^{-2} - 6 \chi^{-1},\ c \hbox{ arbitrary}. \label{eqChazyFamily2} \end{eqnarray} The linearized equation at (\ref{eqChazyOrder0}) \begin{equation} E'(x,u^{(0)}) u^{(1)} \equiv [ \partial_x^3 - 2 u^{(0)} \partial_x^2 + 6 u^{(0)}_x \partial_x - 2 u^{(0)}_{xx}] u^{(1)} = 0, \label{eqChazyLin} \end{equation} has effectively order $N$ (which means that solution (\ref{eqChazyOrder0}) is a particular one, not a singular one), it possesses the two single valued global solutions $\partial_{x_0} u^{(0)}$ and $\partial_c u^{(0)}$, i.e.~$u^{(1)}=\chi^{-3},\chi^{-2}$, and it has only two singular points $\chi=0,\infty$. The point $\chi=0$ is irregular singular of rank two for the non-Fuchsian family (\ref{eqChazyFamily2}) (and regular singular for the Fuchsian family (\ref{eqChazyFamily1}), with Fuchs indices $-4,-3,-2$), and it defines a third {\it formal} solution admitting an essential singularity at $\chi=0$ (ref.~\cite{Ince} chap.~XVII) \begin{equation} \chi \to 0,\ c \not=0:\ u^{(1)}=e^{a / \chi} \chi^{s} w(\chi), \end{equation} in which $(a,s)$ denote constants and $w(\chi)$ a formal Taylor series generically asymptotic, i.e.~with a null radius of convergence. This is not the case here, where computation yields $a=-2c,s=-2,w=1$, thus defining the fundamental set of {\it global} solutions \begin{eqnarray} \forall \chi\ \forall c:\ u^{(1)} & = & \chi^{-2},\ \chi^{-3},\ (e^{-2 c / \chi} - 1 + 2 c \chi^{-1}) \chi^{-2} / (2 c^2) \label{eqChazySGOrder1} \end{eqnarray} and solving the question for the perturbation order $n=1.$ The point $\chi=\infty$ is therefore, in this too simple an example, useless to study. This point is regular singular with Fuchs indices $2,3,4$, and without any more computation one concludes to the {\it local} singlevaluedness since index $4$ cannot generate a logarithm. {\it Remarks}. \begin{itemize} \item Going on with the formalism of Painlev\'e's lemma at higher orders constitutes the rigorous mathematical framework of the local representation of the general solution obtained by Joshi and Kruskal \cite{JoshiKruskal1993} \begin{equation} u= - 6 \chi^{-1} + c \chi^{-2} (1 + z - z^2 / 8 + z^3 / 144 - 7 z^4 / 13824 + O(\varepsilon^5)),\ z=(\varepsilon / c) e^{- 2 c / \chi}. \end{equation} This representation reduces to the one given by Chazy (Taylor series in $1/ \chi$) if one starts from the Fuchsian family (\ref{eqChazyFamily1}). \item The lowering by $M=2$ units of the order of the linearized equation is obtained by the change of function \begin{equation} u^{(1)}=\chi^{-3} v:\ (\partial_x + 3 \chi^{-1} - 2 c \chi^{-2}) v''=0, \end{equation} which yields the third global solution by three quadratures. \end{itemize} The simplified equation (\ref{eqChazy}), which possesses the PP \cite{ChazyThese} and therefore for which no $u^{(n)}$ is multivalued, only shows the method. We now illustrate on other examples the interest of non-Fuchsian families to detect the presence of a movable critical singularity, very often as soon as the first perturbation order. \section{An equation of Bureau ($N=4,M=2$) \label{eqsectionBureau}} \indent The equation (ref.~\cite{BureauMII} p.~79) \begin{equation} E(x,u) \equiv u'''' + 3 u u'' - 4 u'^2 = 0 \end{equation} possesses the global two-parameter solution \cite{CFP1993} \begin{equation} u^{(0)}=c \chi^{-3} - 60 \chi^{-2},\ \chi=x-x_0, \end{equation} for which the linearized equation \begin{equation} E^{(1)} = E'(x,u^{(0)}) u^{(1)} \equiv [ \partial_x^4 + 3 u^{(0)} \partial_x^2 - 8 u^{(0)}_x \partial_x + 3 u^{(0)}_{xx}] u^{(1)} = 0, \label{eqBureauLin} \end{equation} whose only two singular points are $\chi=0$ and $\chi=\infty$, admits the two global single valued solutions $\partial_{x_0} u^{(0)}$ and $\partial_c u^{(0)}$, i.e.~$u^{(1)}=\chi^{-4},\chi^{-3}$, leaving only two other solutions to examine. For $c=0$ it has four global single valued solutions $u^{(1)}=\chi^{-5},\chi^{-4},\chi^{-3},\chi^{18}$, and from now on we assume $c\not=0$. The point $\chi=0$ is singular irregular with rank one, and the two other solutions are non-Fuchsian and {\it formally} given as \begin{equation} \chi \to 0,\ c \not=0:\ u^{(1)}=e^{\pm \sqrt{-12c/ \chi}}\chi^{31/4} (1 + O(\sqrt{\chi})), \end{equation} detecting the presence in (\ref{eqBureauLin}) of an essential singularity at $\chi=0$, but the generically null radius of convergence of the formal series forbids to immediately conclude to the multivaluedness of $u^{(1)}$. The point $\chi=\infty$ is singular regular, with Fuchs indices $-18,3,4,5$. The highest index, $5$, yields a converging series $u^{(1)}=(1/\chi)^{5} w_{5}(1/\chi)$, locally single valued. As for the test for the existence of a series $u^{(1)}=(1/\chi)^{-18} w_{-18}(1/\chi)$, it detects a logarithm arising from the index $3$ when one tries to solve the recursion relation for the twenty-second coefficient of the series $w_{-18}$. One thus concludes here as soon as order one, to be compared with order seven necessary to ref.~\cite{CFP1993}, after a computation practically untractable without a computer. {\it Remark}. The lowering by $M=2$ units of the order of the linearized equation (\ref{eqBureauLin}) is obtained with \begin{equation} u^{(1)}=\chi^{-4} v:\ (\partial_x^2 -16 \chi^{-1} \partial_x +3 c \chi^{-3} - 60 \chi^{-2}) v'' = 0, \end{equation} and it yields the two other solutions in global form \begin{eqnarray} c \not=0:\ v''_1 & = & \chi^{-3} {}_{0} F_{1} (24;-3c/\chi) = \chi^{17/2} J_{23}(\sqrt{12 c/\chi}), \\ v''_2 & = & \chi^{17/2} N_{23}(\sqrt{12 c/\chi}), \end{eqnarray} where the hypergeometric fonction ${}_{0} F_{1} (24;-3c/\chi)$ is single valued and possesses an isolated essential singularity at $\chi=0$, while the fonction $N_{23}$ of Neumann is multivalued because of a $\mathop{\rm Log}\nolimits \chi$ term. {\it Remark}. The value $n=7$ of the perturbation order in $\varepsilon$ needed by ref.~\cite{CFP1993} is the root of the linear equation $n (i_{\rm min}-p) + (i_{\rm max}-p)=-1$, with $p=-2,i_{\rm min}=-5,i_{\rm max}=18$, linking the pole order $p$ in the Fuchsian case $c=0$, the smallest and the greatest Fuchs indices. It expresses the condition for the first occurence of a power $\chi^{-1}$, leading by integration to a logarithm, in the r.h.s.~$R^{(n)}$ of the linear inhomogeneous equation (\ref{eqOrdern}), r.h.s.~created by the nonlinear terms $3 u u'' - 4 u'^2 $. \section{An example in cosmology: Bianchi IX ($N=6,M=4$) \label{sectionBianchi}} \indent The Bianchi IX cosmological model in vacuum \cite{LandauLifshitzTheorieChamps} is ruled by the sixth order system \begin{equation} \sigma^2 (\mathop{\rm Log}\nolimits A)'' = A^2 - (B-C)^2 \hbox{ and cyclically},\ \sigma^2= \pm 1, \label{eqBianchi1} \end{equation} and it does not possess the PP \cite{CGR1994,LMC1994}. Let us prove it with our method. In the neighborhood of the global solution depending on the four arbitrary parameters $(k_1,k_2,\tau_1,\tau_2)$ \cite{Taub} ($x$ is here denoted $\tau$) \begin{equation} A^{(0)}= \sigma {k_1 \over \mathop{\rm sinh}\nolimits k_1 (\tau-\tau_1)},\ B^{(0)}=C^{(0)} = \sigma {k_2^2 \mathop{\rm sinh}\nolimits k_1 (\tau-\tau_1) \over k_1 \mathop{\rm sinh}\nolimits^2 k_2 (\tau-\tau_2)}, \label{eqTaub} \end{equation} the perturbation \begin{eqnarray} & & A= A^{(0)} (1 + \varepsilon A^{(1)} + O(\varepsilon^2)) \hbox{ and cyclically} \end{eqnarray} generates a linearized system whose order is indeed equal to $N=6$ (which proves that solution (\ref{eqTaub}) is a particular solution, not a singular solution), and the lowering by $M=4$ units of its order is obtained by the change of function dictated by the symmetry of the system: $P^{(1)}=B^{(1)}+C^{(1)},M^{(1)}=B^{(1)}-C^{(1)}$ \cite{LMC1994} \begin{eqnarray} & & \sigma^2 A^{(1)''} - 2 A^{(0)^2} A^{(1)} = 0,\ \label{eqTaub1A} \\ & & \sigma^2 P^{(1)''} - 2 A^{(0)} B^{(0)} P^{(1)} = 4 (A^{(0)} B^{(0)} - A^{(0)^2}) A^{(1)},\ \label{eqTaub1P} \\ & & \sigma^2 M^{(1)''} + 2 (A^{(0)} B^{(0)} - 2 B^{(0)^2}) M^{(1)} = 0. \label{eqTaub1M} \end{eqnarray} Indeed, the four single valued global solutions \begin{equation} (A^{(1)},P^{(1)})=\partial_c (\mathop{\rm Log}\nolimits A^{(0)},\mathop{\rm Log}\nolimits (B^{(0)}+C^{(0)})),\ c=k_1,k_2,\tau_1,\tau_2, \end{equation} are those of the equations (\ref{eqTaub1A})--(\ref{eqTaub1P}), \begin{equation} (A^{(1)},P^{(1)} + 2 A^{(1)})= \left\{ \begin{array}{l} ( (\tau - \tau_1) \coth k_1 (\tau - \tau_1) - 1/k_1 , 0 ), \\ ( 0 , (\tau - \tau_2) \coth k_2 (\tau - \tau_2) - 1/k_2 ), \\ ( \coth k_1 (\tau - \tau_1) , 0 ), \\ ( 0 , \coth k_2 (\tau - \tau_2) ), \end{array} \right. \nonumber \end{equation} and there only remains to study the equation (\ref{eqTaub1M}), the singular points of which (modulo the period of $\mathop{\rm sinh}\nolimits$) are $\tau-\tau_2=0$ and $\tau=\infty$. At $\tau-\tau_2=0$, the equation (\ref{eqTaub1M}) genericaly possesses irregular singular points of rank two since the coefficient $B^{(0)^2}$ has there quadruple poles. Its two non-Fuchsian solutions are {\it formally} (ref.~\cite{Ince} chap.XVII) \begin{equation} \tau - \tau_2 \to 0:\ M^{(1)}= e^{\alpha / (\tau - \tau_2)} \sum_{k=0}^{+ \infty} \lambda_k (\tau-\tau_2)^{k+s},\ \lambda_0 \not=0, \end{equation} with \cite{LMC1994} \begin{equation} \alpha= \pm 2 k_1^{-1} \mathop{\rm sinh}\nolimits k_1 (\tau_2 - \tau_1),\ s =1 \mp 2 \mathop{\rm cosh}\nolimits k_1 (\tau_2 - \tau_1). \end{equation} The two generically irrational values for the exponents $s$ allow to conclude only if the Taylor series $\lambda_k (\tau-\tau_2)^{k}$ can be summed, which is the case at least for $k_1=k_2=0$ where the two solutions are globally known \cite{LMC1994}: \begin{eqnarray} k_1=k_2=0:\ & & {\hbox{d}^2 M^{(1)} \over \hbox{d} t^2} + \left({2 \over t^2} - {4 (t-1)^2 \over t^4} \right) M^{(1)}=0,\ t={\tau - \tau_2 \over \tau_1 - \tau_2}, \nonumber \\ & & M^{(1)}= e^{-2/t} t^{-1},\ e^{-2/t} t^{-1} \int^{1/t} z^{-4} e^{4 z} \hbox{d} z, \end{eqnarray} The second solution implies the presence of a logarithmic branch point at $t=0$, or at $t=\infty$ as well. This singularity persists for $(k_1,k_2)\not=(0,0)$ and this proves the absence of the Painlev\'e property for the Bianchi IX model in vacuum. \section{Conclusion \label{sectionConclusion}} \indent The main application of this extension to the Painlev\'e test is the case of a non-Fuchsian family: if the unperturbed solution is known in closed form and if the linearized equation has no fixed singularity at a finite distance, the study of the point at infinity often allows to conclude to non-integrability (in the sense of absence of the PP) much more rapidly than the above mentioned algebraic methods. The present method constitutes an algorithmic extension to the Painlev\'e test. For the two examples given where the Painlev\'e test fails, this failure was already known, and we do not know yet if there exist equations where a failure would only be detectable by the present extension and, of course, by the $\alpha-$method. Nevertheless, the algorithmic cost of the present method is much smaller than the cost of earlier methods, and this is particularly sensitive on the example of Bureau. Moreover, in our two examples, the perturbation order $n=1$ is sufficient to dectect a failure, which the perturbative Fuchsian test detects at the respective orders $n=7$ (Bureau) and $n=2$ (Bianchi IX). Despite the examination of several other equations without the PP and admitting a non-Fuchsian family, we could not find an example requiring an order $n$ greater than one to exhibit a failure of the test. Such a feature may be generic and certainly deserves future investigations. {\it Acknowledgements}. We thank D.~Bessis, A.~Latifi, A.~Mezincescu and A.~Pickering for fruitful discussions.
\section{Introduction} Kinematics of a very thin vortex tube in three-dimensional fluid may be described by the filament equation in the local induction approximation\ \cite{LI1,LI2}. It is formulated as \begin{equation} \pd{\gamma}{t}=\pd{\gamma}{s}\times\dpd{\gamma}{s}, \label{filament} \end{equation} where $\gamma=\gamma\,(t,s)$ denotes the position of the vortex filament in $\ensuremath{\mathsf R}^3$ with $t$ and $s$ being the time and the arc-length parameter respectively. Hasimoto\ \cite{H} introduced a map $h:\gamma\mapsto\psi=\kappa\, \exp [i\int^s\tau(u)du]$, in order to transform the filament equation into the nonlinear Schr\"{o}dinger (NLS) equation for $\psi$. Here $\kappa$ and $\tau$ respectively denote the curvature and the torsion along $\gamma$. Since the integrability of the NLS equation was well known, the filament equation was naturally expected to be integrable. Mardson and Weinstein\ \cite{MW} first described the filament equation as a Hamiltonian equation with the Hamiltonian simply being the length $\ell\,[\gamma]$ of the vortex filament. Later Langer and Perline\ \cite{LP} used this Hamiltonian structure to prove the existence of an infinite sequence of constants of motion in involution, and studied the evolution of the vortex filaments in connection with the solitons in the NLS equation. With this concern in mind, we have investigated the filament equation in a curved three-manifold $M$. Although Langer and Perline have limited $M$ to $\ensuremath{\mathsf R}^3$, we find an analogous integrable hierarchy in the case of constant curvature. We further study the classical partition function for the vortex filaments \begin{equation} Z(\beta)=\int_{\Gamma}\, e^{-\beta\,\ell\,[\gamma]}\,{\cal D} \gamma{}. \label{partition} \end{equation} It is not clear if the Duistermaat-Heckman formula\ \cite{DH} applies to this case, because our phase space $\Gamma$ is neither finite dimensional nor compact, and furthermore because the Hamiltonian flow may not be periodic. But the perturbative calculation in our mode reveals that the loop corrections to the formula vanish up to the 3-loop. \section{Integrability} We begin this section by describing a symplectic structure for the vortex filament in a three-manifold $M$ equipped with a Riemann metric $g$. Everything is considered in the smooth category for simplicity. Let $\Gamma$ be the space of vortex filaments with fixed end points $p$ and $q$; $\Gamma$ is the quotient space of $\{\gamma:[0,1]\rightarrow M \mid\gamma(0)=p,\gamma(1)=q\}$ with the reparametrization of $\gamma$. Hereafter $\gamma$ denotes the representative for which the parameter $x\in[0,1]$ is a multiple of the arc-length $s$, namely \begin{equation} \frac{ds}{dx}=\norm{\frac{d\gamma}{dx}}= \sqrt{(\frac{d\gamma}{dx},\frac{d\gamma}{dx})} \end{equation} is independent of $x$. Here $(\ \,,\ )$ denotes the inner product on the tangent space $T_{\gamma\,(x)}M$. One can identify the tangent space $T_{\gamma}\Gamma$ with the subspace of $\Gamma(\gamma^{\ast}TM)$, and expand $X\in\Gamma(\gamma^{\ast}TM)$ in the Frenet-Serret frame along $\gamma$ such that \begin{equation} X=f\,\ensuremath{\mathsf T}+g\,\ensuremath{\mathsf N}+h\,\ensuremath{\mathsf B}, \end{equation} where $\ensuremath{\mathsf T}$ is the unit tangent vector to $\gamma$, $\ensuremath{\mathsf N}$ is the unit normal vector and $\ensuremath{\mathsf B}$ is the unit binormal vector. Let $\ell\,[\gamma]$ be the length of $\gamma$, so that $s=\ell\,[\gamma]\,x$. The Frenet-Serret equations are \begin{equation} \nabla_{s}\ensuremath{\mathsf T}=\kappa\,\ensuremath{\mathsf N},\ \ \ \ \nabla_{s}\ensuremath{\mathsf N}=-\kappa\,\ensuremath{\mathsf T}+\tau\,\ensuremath{\mathsf B},\ \ \ \ \nabla_{s}\ensuremath{\mathsf B}=-\tau\,\ensuremath{\mathsf N}, \label{FS} \end{equation} with $\nabla$ being the connection on $\gamma^{\ast}TM$ induced by the Levi-Civita connection on $TM$. Let $\wp$ be the projection from $\Gamma(\gamma^{\ast}TM)$ to $T_{\gamma}\Gamma$, then one can show that the tangent component of $v=\wp\,(X)\in T_{\gamma}\Gamma$ satisfies \begin{equation} \frac{d}{dx}\,v_{\ensuremath{\mathsf T}}=\ell^{-1}(\nabla_{x}v,\frac{d\gamma}{dx})+ \ell\,\kappa\,v_{\ensuremath{\mathsf N}}, \end{equation} and $(\nabla_{x}v,d\gamma/dx)$ is a constant. Fixing this constant by the boundary conditions $X(0)=X(1)=0$, one obtains \begin{equation} \wp\,(X)=v=\ell\,(\int^{x}_{0}\kappa\,v_{\ensuremath{\mathsf N}}dx - x\,\int^{1}_{0}\kappa\,v_{\ensuremath{\mathsf N}}dx)\,\ensuremath{\mathsf T} +v_{\ensuremath{\mathsf N}}\ensuremath{\mathsf N}+v_{\ensuremath{\mathsf B}}\ensuremath{\mathsf B}\ . \end{equation} Geometrical structures on $\Gamma$ were first studied by Marsden and Weinstein\ \cite{MW} for the vortex filament in $\ensuremath{\mathsf R}^3$, and generalized to the loop space for a three-manifold $M$ by Brylinski\ \cite{B}. It is straightforward to find those for the vortex filament in $M$. {\parindent=0pt i) Complex structure} For the tangent vector $v\in T_{\gamma}\Gamma$, $J$ generates the 90-degree rotation \begin{equation} J(v)=-\,\wp\,(\ensuremath{\mathsf T}\times v)\ ,\ \ \ J^2=-1. \end{equation} Choosing $(v_{\ensuremath{\mathsf N}},v_{\ensuremath{\mathsf B}})$ as coordinates for $T_{\gamma}\Gamma$, we find that $J$ corresponds to the multiplication by $i$ for the complex function $v_{\ensuremath{\mathsf N}}(x)+i\,v_{\ensuremath{\mathsf B}}(x)$. Hence $J$ induces a complex structure on $\Gamma$. {\parindent=0pt ii) Riemann structure} The Riemann structure on $\Gamma$ is simply defined by \begin{equation} \ip{u}{v}_{\Gamma}=\ell\,\int_{0}^{1}(u_{\ensuremath{\mathsf N}}v_{\ensuremath{\mathsf N}}+u_{\ensuremath{\mathsf B}}v_{\ensuremath{\mathsf B}})\,dx \end{equation} for $u,v\in T_{\gamma}\Gamma$, and satisfies the hermitian condition \begin{equation} \ip{u}{v}_{\Gamma}=\ip{J(u)}{J(v)}_{\Gamma}. \end{equation} Note that even though $\ip{\ \,}{\ }_{\Gamma}$ ignores the $\ensuremath{\mathsf T}$-components, it is non-degenerate. {\parindent=0pt iii) Symplectic structure} The volume form $\nu$ on $M$ associated with the Riemann metric $g$ provides the symplectic structure on $\Gamma$, namely \begin{equation} \omega(u,v)=\int_0^1\nu(\frac{d\gamma}{dx},u,v)\,dx. \end{equation} Using the Frenet-Serret frame, one can rewrite this as \begin{equation} \omega(u,v)=\ell\,\int_0^1(u_{\ensuremath{\mathsf N}}v_{\ensuremath{\mathsf B}}-u_{\ensuremath{\mathsf B}}v_{\ensuremath{\mathsf N}})\,dx, \end{equation} which is equivalent to the one constructed from the above two structures \begin{equation} \omega(u,v)=\ip{u}{J(v)}_{\Gamma}. \end{equation} Having set out the basic structures, we now turn to the Hamiltonian flows for the vortex filament. Let $\ell\,:\Gamma\mapsto\ensuremath{\mathsf R}$ be a smooth Hamiltonian function, then the Hamiltonian vector field $X_{\ell}$ has the form \begin{equation} X_{\ell}=J(\ensuremath{\mathrm{grad}\,}\ell). \end{equation} Choosing $i_{X_{\ell}}\,\omega=d\,\ell$ and putting $v=d\gamma_t/dt\mid_{t=0}$, we get \begin{eqnarray} v\,\ell\,[\gamma]&=&\frac{d}{dt}\left.\int_0^1 \sqrt{\left(\pd{\gamma_t}{x}, \pd{\gamma_t}{x}\right)}dx\right|_{t=0},\nonumber\\ &=&\frac{1}{\ell\,[\gamma]}\int_0^1 \left(\nabla_xv,\frac{d\gamma}{dx}\right)dx,\\ &=&-\ell\,[\gamma]\,\int_0^1\left(v,\kappa\,\ensuremath{\mathsf N}\right)dx.\nonumber \end{eqnarray} $\ensuremath{\mathrm{grad}\,}\ell=-\wp(\kappa\,\ensuremath{\mathsf N})$ follows, and therefore \begin{equation} X_{\ell}=\kappa\,\ensuremath{\mathsf B}. \end{equation} This yields a natural generalization of the filament equation in $M$\ \cite{K} \begin{equation} \pd{\gamma}{t}=\kappa\,\ensuremath{\mathsf B} =\ell^{-3}\,\pd{\gamma}{x}\times\nabla_{x}\pd{\gamma}{x}. \end{equation} The evolution equations for $\kappa$ and $\tau$ are the followings \begin{eqnarray} \pd{\kappa}{t}&=&\kappa\,\ensuremath{\mathrm{Ric}\,}(\ensuremath{\mathsf B},\ensuremath{\mathsf N})-\ell^{-1} (2\tau\pd{\kappa}{x}+\kappa\pd{\tau}{x}),\\ \pd{\tau}{t}&=&\tau\,\ensuremath{\mathrm{Ric}\,}(\ensuremath{\mathsf T},\ensuremath{\mathsf N})+\ell^{-1} \pd{}{x}(\frac{1}{2}\kappa^2+\ell^{-2}\kappa^{-1}\dpd{\kappa}{x} -\tau^{2}+\rho(\ensuremath{\mathsf T},\ensuremath{\mathsf B})), \end{eqnarray} where $\ensuremath{\mathrm{Ric}\,}$ and $\rho$ denote the Ricci tensor and the sectional curvature on $M$ respectively. In the case of constant curvature, these equations take simpler forms \begin{eqnarray} \pd{\kappa}{t}&=&-\ell^{-1} (2\tau\pd{\kappa}{x}+ \kappa\pd{\tau}{x}),\label{evol1}\\ \pd{\tau}{t}&=&\ell^{-1}\pd{}{x} (\frac{1}{2}\kappa^2+\ell^{-2}\kappa^{-1}\dpd{\kappa}{x} -\tau^{2}),\label{evol2} \end{eqnarray} which turn out to be identical with the equations appeared in\ \cite{H}. Hence we can get the following proposition for a three-manifold with constant curvature. \vskip 8pt {\parindent=0pt {\bf Proposition}} \begin{itemize} \item[(a)] The filament equation is transformed into the NLS equation by the Hasimoto map. \item[(b)] There is an infinite sequence of constants of motion. \item[(c)] These constants are in involution. \end{itemize} \vskip 8pt {\parindent=0pt {\bf Proof}} We assume that $\kappa$, $\tau$ and their derivatives of arbitrary order vanish at the boundaries. Then it is straightforward to prove (a) and (b) due to the evolution equations\ (\ref{evol1}) and (\ref{evol2}). Using the explicit form of the Hamiltonian vector fields $X_n$ (see Remark (2)), we can confirm the commutativity $\omega(X_n,X_m)=0$ for any $n$ and $m$ with the help of\ \cite{LP}, and consequently prove (c). \vskip 8pt {\parindent=0pt {\bf Remarks}} \begin{itemize} \item[(1)] The constants of motion are as follows\ \cite{LP}: \begin{eqnarray} I_{-2}[\gamma]&=&\ell\,[\gamma], \ \ \ I_{-1}[\gamma]= \ell\,\int_{0}^{1}\tau\,dx,\\ I_{n}[\gamma]&=&\tilde I_{n}\circ h[\gamma]\ \ \ (n=0,1,2,\ldots), \nonumber \end{eqnarray} where $h$ is the Hasimoto map $h[\gamma]=\kappa\,\exp[i\ell\,[\gamma]\int_0^x\tau\,dx]$, and $\tilde I_{n}$'s are the constants of motion in the NLS equation\ \cite{FT} given by \begin{equation} \tilde I_{n}[\psi]=\ell\,\int_{0}^{1} \ensuremath{\frac{1}{2}}\,\bar\psi\,\tilde J_{n}(\psi,\bar\psi)\,dx, \end{equation} and \begin{equation} \tilde J_{0}=\psi,\ \ \ \tilde J_{n+1}=-i\,\frac{d}{ds}\tilde J_{n}- \frac{1}{4}\,\bar\psi\sum_{k=1}^{n}\tilde J_{k-1}\,\tilde J_{n-k}. \end{equation} \item[(2)] We find mutually commuting Hamiltonian vector fields $X_{n}$ for $I_{n}[\gamma]$: \begin{eqnarray} X_{-2}&=&\kappa\,\ensuremath{\mathsf B}, \ \ \ X_{-1}=R\,X_{-2},\\ X_{n}&=&R^{n+2}X_{-2}-c\,R^{n}X_{-2}\ \ \ (n=0,1,2,\ldots), \nonumber \end{eqnarray} where $c$ denotes the constant curvature and $R$ the ``recursion operator" defined by \begin{equation} R(v)=-\ell^{-1}\,\wp(\ensuremath{\mathsf T}\times\nabla_{x}v) \end{equation} for $v\in T_{\gamma}\Gamma$. $R$ coincides with the one appeared in\ \cite{LP} when we restrict $v$ to the Hamiltonian vector fields. \item[(3)] Langer and Perline interpreted the Hasimoto map as a Poisson map between the Poisson structure on the space of the vertex filaments and the ``forth" Poisson structure on the space of the NLS fields. We have found no such correspondence in our model, because the deformation of the vortex filament also changes its length $\ell$. In the case of\ \cite{LP}, however, the vortex filament extends boundlessly, so that the arc-length parameter is simply a parameter and does not change under the deformation. A different approach to the integrability of the vortex filaments has been investigated in\ \cite{S} recently. \end{itemize} The filament equation belongs to an infinite hierarchy of Hamiltonian systems $\{\partial\gamma/\partial t_n=X_n\mid n=-2,-1,0,\ldots\}$, and all Hamiltonian flows in this hierarchy are transformed into those in the NLS hierarchy. In fact, the differential of $h$ yields \begin{equation} dh:X_n\longmapsto\tilde{X}_{n+4}-2c\,\tilde{X}_{n+2}+c^2\, \tilde{X}_{n}\ \ \ \pmod{i\psi}, \end{equation} where $\tilde{X}_{-2}=\tilde{X}_{-1}=0$, and $\tilde{X}_{n}\ (n=0,1,2,\ldots)$ are the Hamiltonian vector fields associated with $\tilde{I}_n[\psi]$, {\it i.e.}, $\tilde{X}_{n}=-i\,\ensuremath{\mathrm{grad}\,}\tilde{I}_{n}$; first two are \begin{eqnarray} dh(X_{-2})&=&i\,(\frac{d^2\psi}{ds^2}+ \frac{1}{2}\mid\psi\mid^{2}\psi),\\ dh(X_{-1})&=&\frac{d^3\psi}{ds^3}+ \frac{3}{2}\mid\psi\mid^{2}\frac{d\psi}{ds}+ 2c\,\frac{d\psi}{ds}. \end{eqnarray} \section{Classical partition function} In this section we evaluate the classical partition function\ (\ref{partition}) with ${\cal D}\gamma$ being the symplectic volume form on $\Gamma$. The stationary phase method provides an asymptotic expansion for $Z(\beta)$ as $\beta\mapsto\infty$, such that \begin{equation} Z(\beta)=\sum_{\ensuremath{\mathrm{grad}\,}\ell\,[\gamma]\,=0}Z_{\mathrm WKB}[\gamma,\beta]\, (1+\frac{a_1[\gamma]}{\beta}+\frac{a_2[\gamma]}{\beta^2}+\cdots). \label{asymptotic} \end{equation} The exactness of the stationary phase (WKB) approximation has been of interest due to the Duistermaat-Heckman formula\ \cite{DH}, where they have shown that if $\Gamma$ is a compact symplectic manifold and $\ell$ is a periodic Hamiltonian with isolated critical points, WKB approximation becomes exact for\ (\ref{partition}), {\it i.e.}, the asymptotic expansion terminates at $Z_{\mathrm WKB}$. In more general arguments presented in\ \cite{A}, the fixed points are not necessarily isolated, and it is not mandatory to consider the circle action alone according to the analogous results obtained for higher dimensional tori. For the infinite dimensional symplectic manifolds, the WKB exactness has not been proved rigorously, but a ``proper" version of WKB approximation should yield a reliable result for a large class of integrable models\ \cite{N1,N2,N3,N4}. With this notion in mind, we present the explicit calculation of the asymptotic expansion\ (\ref{asymptotic}). For simplicity, we will assume the followings: \begin{itemize} \item[(1)] $M$ is a three-manifold with a constant curvature $c$, so that the filament equation is integrable in the sense of Proposition. \item[(2)] Two points $p$ and $q$ on $M$ are not conjugate. Consequently, the Hamiltonian $\ell$ is a Morse function on $\Gamma$, {\it i.e.}, critical points are the geodesics on $M$ connecting $p$ and $q$, and further the Hessian operator $H_{\gamma}$ at each geodesic $\gamma$ is a non-degenerate Jacobi operator \begin{equation} H_{\gamma}=-\nabla_{x}\nabla_{x}-c\,\ell[\gamma]^2. \end{equation} \end{itemize} Let us first expand the Hamiltonian $\ell$ around a geodesic $\gamma$. As we can see in (\ref{FS}), the curvature along the geodesic vanishes identically, and $\xi(x)\in T_{\gamma(x)}M$ thus satisfies the condition $(\xi,\ensuremath{\mathsf T})=0$. Using an infinitesimal deformation of $\gamma$ generated by the exponential map $\gamma_s(x)=\exp_{\gamma(x)}[s\,\ell\,\xi(x)]$, we can find the expansion \begin{eqnarray} \ell\,[\gamma_s]&=&\sum_{n=0}^{\infty}\frac{s^n}{n!} \left.\frac{d^n}{ds^n}\ell\,[\gamma_s]\right|_{s=0},\\ &=&\ell\,[\gamma]\sum_{n=0}^{\infty}\frac{s^{2n}}{(2n)!} \int_0^1 W_{2n}(\xi)dx. \end{eqnarray} Here the integrand $W_{2n}$ is given by the Bell Polynomial $Y_m$\ \cite{Bell}, namely \begin{equation} W_{2n}(\xi)=Y_{2n}(f_{1},\ldots,f_{2n};g_1(\xi),\ldots,g_{2n}(\xi)), \end{equation} with \begin{eqnarray} f_m&=&(-)^{m-1}\frac{(2m-3)!!}{2^m},\nonumber\\ g_2(\xi)&=&2\,(\nabla_x\xi,\nabla_x\xi)-2c\,\ell^2(\xi,\xi), \nonumber\\ g_{2m}(\xi)&=&(-)^{m}2^{2m-1}\,\{ (c\,\ell^2)^{m}(\xi,\xi)^{m}\label{vertices}\\ &+&(c\,\ell^2)^{m-1}(\xi,\xi)^{m-2} [(\nabla_x\xi,\xi)^2-(\xi,\xi)(\nabla_x\xi,\nabla_x\xi)]\} \ \ \ (m\geq 2),\nonumber\\ g_{2m+1}(\xi)&=&0.\nonumber \end{eqnarray} First few are given by \begin{equation} \begin{array}{lcl} W_0=1, & \ \ \ \ & W_2=f_1\,g_2,\\ &&\\ W_4=f_1\,g_4+3\,f_2\,g_2^2, & \ \ \ \ & W_6=f_1\,g_6+15\,f_2\,g_4\,g_2+15\,f_3\,g_2^3.\\ \end{array} \end{equation} Now let us evaluate the WKB partition function \begin{eqnarray} Z_{\mathrm WKB}[\gamma,\beta]&=&e^{-\beta\ell[\gamma]} \int{\cal D}\,\xi \exp\left[-\frac{\beta\ell[\gamma]}{2}\int_0^1W_2(\xi)dx\right], \label{WKB}\\ \int_0^1W_2(\xi)dx&=&\ip{\xi}{H_{\gamma}(\xi)}_{\Gamma}. \end{eqnarray} Using the zeta-function regularization technique, we can perform the infinite dimensional integral in\ (\ref{WKB}), and obtain\ \cite{AS} \begin{equation} Z_{\mathrm WKB}[\gamma,\beta]= e^{-\beta\,\ell[\gamma]\pm\frac{\pi}{4} i(\eta_{H}(0)-\zeta_{H}(0))} e^{\frac{1}{2}\zeta_{H}'(0)}\, (\beta\,\ell[\gamma])^{-\frac{1}{2}\zeta_{H}(0)}, \end{equation} with $\eta_{H}(z)$ and $\zeta_{H}(z)\ (z\in\ensuremath{\mathsf C})$ being eta and zeta functions associated with the Hessian operator $H_{\gamma}$ respectively. Evaluating these functions for $\gamma$ with the Morse index $\mu(\gamma)$, we find \begin{equation} \eta_{H}(0)=-1-2\,\mu(\gamma), \ \ \ \zeta_{H}(0)=-1, \ \ \ \zeta_{H}'(0)=2\ln\left|\frac{\sqrt{c}\,\ell[\gamma]} {2\sin(\sqrt{c}\,\ell[\gamma])}\right|, \end{equation} and eventually this gives us an explicit expression \begin{equation} Z_{\mathrm WKB}[\gamma,\beta]=\frac{1}{2}\,e^{-\beta\,\ell[\gamma]} \sqrt{\beta\,\ell[\gamma]} \left|\frac{\sqrt{c}\,\ell[\gamma]}{\sin(\sqrt{c}\,\ell[\gamma])} \right| e^{\mp\frac{\pi}{2}i\,\mu(\gamma)}. \end{equation} Since $\mu(\gamma)$ is an even integer, the last factor contains no ambiguities. We now proceed to the higher-order calculation. It is convenient to choose an orthogonal frame $\{e_1,e_2\}$ along $\gamma$ such that \begin{equation} \nabla_x\,e_i=0,\ \ \ (\ensuremath{\mathsf T},e_i)=0 \ \ \ \mathrm {for} \ i=1,2. \end{equation} In this frame, the kernel of the Jacobi operator $H_{\gamma}$ becomes diagonal, and both of the diagonal elements are identical to the Dirichlet Green function \begin{equation} G(x,x')=2\sum_{n=1}^{\infty}\frac{\sin(n\pi x)\sin(n\pi x')} {(n\pi)^2-\lambda}, \end{equation} with $\lambda=c\,\ell^2$. The 2-loop amplitude $a_1=-\beta^2\,\langle W_4/4!\rangle$ consists of four diagrams depicted in Fig.\ 1, and those are respectively \begin{eqnarray} &(a)\hskip 30pt& \lambda^2\,\int_0^1\,G(x)^2= \frac{\lambda}{8}-\frac{3}{8}\sqrt{\lambda}\,X+\frac{3}{8}\lambda\, X^2, \nonumber\\ &(b)\hskip 30pt& \lambda\,\int_0^1\,G'(x)^2= \frac{\lambda}{8}-\frac{1}{8}\sqrt{\lambda}\,X+\frac{1}{8}\lambda\, X^2, \nonumber\\ &(c)\hskip 30pt& \lambda\,\int_0^1\,G(x)\,G''(x)= -\frac{\lambda}{8}-\frac{1}{8}\sqrt{\lambda}\,X+\frac{1}{8}\lambda\, X^2,\nonumber\\ &(d)\hskip 30pt& \,\int_0^1\,G''(x)^2=\lambda^2\,\int_0^1\,G(x)^2, \nonumber \end{eqnarray} where $X=\cot\sqrt{\lambda}$, $G(x)=G(x,x')\mid_{x=x'}$, $G'(x)=(\partial/\partial x)\,G(x,x')\mid_{x=x'}$ and $G''(x)=(\partial^2/\partial x\,\partial x')\,G(x,x')\mid_{x=x'}$. While $G(x)$ and $G'(x)$ are convergent, $G''(x)$ diverges at the boundaries, thus we have found (c) and (d) by executing the $x$-integration first and then by regularizing the infinite $n$-summation in terms of the following analytic continuations: \begin{eqnarray} \sum_{n=1}^{\infty}\frac{1}{(n^2+a^2)^s}&=& -\ensuremath{\frac{1}{2}} a^{-2s}+\frac{\pi^{\ensuremath{\frac{1}{2}}}}{2}\, \frac{\Gamma(s-\ensuremath{\frac{1}{2}})}{\Gamma(s)}\,a^{-2s+1}\nonumber\\ &+&2\,\frac{\pi^{\ensuremath{\frac{1}{2}}}}{\Gamma(s)}\sum_{n=1}^{\infty} \left(\frac{\pi n}{a}\right)^{s-\ensuremath{\frac{1}{2}}}\,K_{s-\ensuremath{\frac{1}{2}}}(2\pi na),\\ \sum_{n=1}^{\infty}\frac{n^2}{(n^2+a^2)^s}&=& \frac{\pi^{\ensuremath{\frac{1}{2}}}}{4}\,\frac{\Gamma(s-\frac{3}{2})}{\Gamma(s)}\, a^{-2s+3}\nonumber\\ &+&\frac{\pi^{\ensuremath{\frac{1}{2}}}}{\Gamma(s)}\,\sum_{n=1}^{\infty} \left(\frac{\pi n}{a}\right)^{s-\frac{3}{2}}\, K_{s-\frac{3}{2}}(2\pi na)\\ &-&2\,\frac{\pi^{\frac{5}{2}}}{\Gamma(s)}\,\sum_{n=1}^{\infty} \,n^2\,\left(\frac{\pi n}{a}\right)^{s-\frac{5}{2}}\, K_{s-\frac{5}{2}}(2\pi na),\nonumber\\ \sum_{n=1}^{\infty}\frac{n^4}{(n^2+a^2)^s}&=& -\frac{3}{8}\,\pi^{\ensuremath{\frac{1}{2}}}\, \frac{\Gamma(s-\frac{5}{2})}{\Gamma(s)}\, a^{-s+\frac{5}{2}}\nonumber\\ &+&\frac{3}{2}\,\frac{\pi^{\ensuremath{\frac{1}{2}}}}{\Gamma(s)}\, \sum_{n=1}^{\infty}\left(\frac{\pi n}{a}\right)^{s-\frac{5}{2}}\, K_{s-\frac{5}{2}}(2\pi na)\nonumber\\ &-&6\,\frac{\pi^{\frac{5}{2}}}{\Gamma(s)}\,\sum_{n=1}^{\infty} \,n^2\,\left(\frac{\pi n}{a}\right)^{s-\frac{7}{2}}\, K_{s-\frac{7}{2}}(2\pi na)\\ &+&2\,\frac{\pi^{\frac{9}{2}}}{\Gamma(s)}\,\sum_{n=1}^{\infty} \,n^4\,\left(\frac{\pi n}{a}\right)^{s-\frac{9}{2}}\, K_{s-\frac{9}{2}}(2\pi na),\nonumber \end{eqnarray} where $K_{\nu}(z)$ is the modified Bessel function. Multiplying (a) through (d) with the weights of the diagrams, we conclude that the 2-loop amplitude vanishes. Beyond the 2-loop, however, we ought to generalize the analytic continuation for a multiple infinite summation. One might think that applying the analytic continuation method directly to the Green function, we could regularize the Green function, and thereby making all loop amplitude finite. This is certainly true, but regularizing the Green function in this way, we also eliminate the necessarily singularity at $x=x'$, and obtain non-vanishing 2-loop amplitude as a result. We may avoid this difficulty by treating $G(x,x')$ as a distribution w.r.t.\ $x$. Let us first examine this on the 2-loop and check if the amplitude vanishes. Since $G(x,x')$ may naturally be extended periodically (period 2) to $\ensuremath{\mathsf R}$ as a function of $x$, one can redefine it as a distribution $\tilde G(x,x')$ such that \begin{eqnarray} \tilde G(x,x')&=&-\frac{1}{\sqrt{\lambda}\sin\sqrt{\lambda}} \sum_{n\in\ensuremath{\mathsf Z}}\, \left\{\sin[\sqrt{\lambda}\,(x-2n)]\,\sin[\sqrt{\lambda}\, (x'-1)]\,H(x;2n,x'+2n)\right.\nonumber\\ &\ &\hskip 30pt +\left.\sin[\sqrt{\lambda}\,x']\,\sin[\sqrt{\lambda}\,(x-2n-1)]\, H(x;x'+2n,2n+1)\right\}\label{Green}\\ &+&\frac{1}{\sqrt{\lambda}\sin\sqrt{\lambda}} \sum_{n\in\ensuremath{\mathsf Z}}\,\left\{\,x\rightarrow -x\,\right\},\nonumber \end{eqnarray} where $H(x;a,b)$ denotes the characteristic function for the interval $[a,b]\subset\ensuremath{\mathsf R}$. Similarly $\tilde G(x)$ may also be extended periodically (period 1) to $\ensuremath{\mathsf R}$ \begin{equation} \tilde G(x)=-\frac{1}{\sqrt{\lambda}\sin\sqrt{\lambda}} \sum_{n\in\ensuremath{\mathsf Z}}\,\left\{\sin[\sqrt{\lambda}\,(x-n)]\, \sin[\sqrt{\lambda}\,(x-n-1)]\,H(x;n,n+1)\right\}. \label{loop} \end{equation} Using the periodic delta function $\delta(x;n)$ ($n$ is the period), we may evaluate the second derivative \begin{eqnarray} \frac{\partial^2}{\partial x\,\partial x'}\,\tilde G(x,x')&=& -\frac{\sqrt{\lambda}}{\sin\sqrt{\lambda}}\sum_{n\in\ensuremath{\mathsf Z}}\, \left\{\cos[\sqrt{\lambda}\,(x-2n)]\,\cos[\sqrt{\lambda}\,(x'-1)]\, H(x;2n,x'+2n)\right.\nonumber\\ &\ &\hskip 30pt +\left.\cos[\sqrt{\lambda}\,x']\,\cos[\sqrt{\lambda}\,(x-2n-1)]\, H(x;x'+2n,2n+1)\right\}\\ &-&\!\!\!\frac{\sqrt{\lambda}}{\sin\sqrt{\lambda}} \sum_{n\in\ensuremath{\mathsf Z}}\,\left\{\,x\rightarrow -x\,\right\}+ \delta(x-x';2)+\delta(x+x';2),\nonumber \end{eqnarray} and similarly \begin{equation} \tilde G''(x)=-\lambda\,\tilde G(x)-\sqrt{\lambda}\,\cot\sqrt{\lambda}+ \frac{1}{2}\,\delta(x;1). \end{equation} The delta function appears only in $\tilde G''(x)$, and we confirm the vanishing of the 2-loop amplitude by using \begin{equation} \int_0^1dx\,\delta(x;1)^2=\delta(0;1)=0, \end{equation} which is consistent with the $\zeta$-function regularization because of $\delta(0;1)=1+2\,\zeta(-1)$. The 3-loop amplitude $a_2=\beta^3\,\langle\beta (W_4/4!)^2/2-W_6/6!\rangle$ consists of 30 diagrams depicted in Fig.\ 2. Evaluating them by means of\ (\ref{Green}), (\ref{loop}) and $\delta(0;1)=0$, we find that 29 diagrams contain no ambiguities due to the the integration formulae \begin{eqnarray} \int_0^1dx\int_0^xdy\,\delta(x;1)\,\delta(y;1)\,F(x,y) &=&\frac{1}{8}\,F(0,0)+\frac{1}{4}\,F(1,0)+\frac{1}{8}\,F(1,1),\\ \int_0^1dx\int_0^xdy\,[\delta(x-y;2)+\delta(x+y;2)]&\ &\hskip -24pt [\delta(x;1)+\delta(y;1)]\,F(x,y)\nonumber\\ &=&\frac{1}{2}\,F(0,0)+\frac{1}{2}\,F(1,1),\label{pq}\\ \int_0^1dx\int_0^xdy\,[\delta(x-y;2)+\delta(x+y;2)]^2\,F(x,y)&=& \frac{1}{8}\,F(0,0)+\frac{1}{8}\,F(1,1). \end{eqnarray} Here the last equality follows from $\delta(0;2)=0$. Yet, in the diagram whose weight is $-480$, we encounter an ambiguous integral \begin{equation} \int_0^1dx\int_0^xdy\,[\delta(x-y;2)+\delta(x+y;2)]\delta(x;1)\, F(x,y)=p\,F(0,0)+q\,F(1,1), \end{equation} where $p+q=1/2$ as is shown in\ (\ref{pq}), but $p$ or $q$ alone cannot be determined unless we specify the regularization of the delta function. If we were able to define the analytic continuation of the infinite double sum, this ambiguity would not appear, but we have no choice at our hand other than putting $q=1/16$, and obtain the vanishing 3-loop amplitude as a result. Ambiguities appearing in higher loops are inevitable, because they relate to the regularization ambiguity of the integration measure ${\cal D}\gamma$, which has never been defined rigorously in the first place. Both methods we have presented here reveal that the degree of ambiguity gets larger as the order of loops increases. In the analytic continuation method, ambiguity arises from the variety of the analytic continuation applicable to the multiple infinite summation, whereas in the distribution method, the delta-function integration, particularly the finite part of the boundary contribution, is the source of the ambiguity. Nevertheless our lower order calculations suggest that by regularizing ${\cal D}\gamma$ order by order, one can eliminate all higher loop corrections, and thereby preserving the Duistermaat-Heckman formula. The symplectic structure has been studied thoroughly in compact finite dimensional manifolds, but little is known for the infinite dimensional ones, which include most of the integrable hierarchies. This is exactly the place where the physical interests are, and the Duistermaat-Heckman formula would throw a new light over the integrable hierarchies as we have caught a glimpse of it here. \vskip 12pt The authors would like to thank Dr.\ N.\ Sasaki for helpful discussions.
\section{Introduction} \label{sec1} The observation in 1984 by Shechtman et al\cite{Shechtman} of a sharp diffraction pattern in a AlMn alloy with the symmetry of the icosahedron, opened a new field in condensed matter physics. Ever since experimental evidence of other materials having sharp diffraction patterns with symmetries forbidden by classical crystallography has continued to grow. The name quasicrystals has been coined to represent systems with perfect order but without periodicity, i.e. quasiperiodic systems. Since the nonperiodic three-dimensional ($3D$) Penrose tiling \cite{Elser} has a diffraction pattern closely similar to that of icosahedral alloys, it has been extensively studied to account for icosahedral point symmetry and also because of its relative simplicity. The $3D$ Penrose tiling is usually constructed by projection from a $6D$ simple cubic lattice. The projection is performed by first defining an acceptance domain in the $3D$ complementary space in order to select what points of the $6D$ simple cubic lattice are effectively projected to form the $3D$ quasilattice. In the present investigation the $3D$ Penrose tiling has been generated by a special choice of the shape and size of the acceptance domain as described by Elser\cite{Elser}. The last step to model a quasicrystal concerns the decoration i.e., the location of the lattice points forming the quasicrystal and a choice for the pair interaction potential. Of the many questions about quasicrystals, one concerns the stability of these phases. The first theoretical approaches to such a question\cite{mermin,kalu} were based upon the Landau theory of crystallization where the free energy is expanded in powers of an order parameter related to density waves with icosahedral symmetry. As a main result it was shown that multi-component systems are required to achieve stability, a fact that agrees with the experimental findings\cite{mermin}. On the other hand and from a more microscopic viewpoint, the analysis of the stability of quasicrystals may be rather difficult on general grounds, but an important simplification occurs if a hard-sphere pair potential is assumed. Indeed, a simple calculation of the maximum packing fraction, i.e. the fraction of the total volume occupied by the spheres, provides an important criterion of the stability chances. For instance, a $3D$ Penrose tiling with all vertices occupied by identical hard spheres leads to a fluid-like packing fraction and thus it must certainly be discarded as a model of a quasicrystal. Moreover, the interest of considering hard-sphere quasicrystals goes far beyond simplicity. Thus, numerical studies\cite{Roth} have shown that the crucial criterion for the quasicrystal stability with more realistic, e.g., Lennard-Jones, interactions is the packing fraction of the quasicrystalline hard-sphere decoration. There are basically two ways of improving the poor packing fraction of the above fully occupied Penrose tiling. These options are to change the decoration\cite{Henley} or the acceptance domain\cite{Oguey}. Both procedures give approximately the same optimal packing fraction $\simeq 0.63$, a value that now indeed justifies a further stability study. Recently, McCarley and Ashcroft\cite{Ash} have studied the hard-sphere quasicrystal using a modification of the acceptance domain\cite{Oguey}, to obtain from the modified weighted density approximation a metastable quasicrystal with respect to the crystalline and fluid phases. Their method is entirely formulated in the $6D$ reciprocal space which avoids direct summations over the quasilattice. However, a drawback of this method is that a truncation of the sum in the $6D$ reciprocal lattice is needed, the induced estimate error in the free energy per particle of the quasilattice being $\simeq 2\%$. In the present paper we will consider the above optimal hard-sphere decoration of the Penrose tiling. Our treatment is based on the generalized effective liquid approximation\cite{Lutsko} which has been previously applied to perfect hard spheres and hard disks crystals yielding very accurate results as compared to the simulation data. The quasilattice sums are calculated in the $3D$ real space using a method which substantially improves convergence errors in comparison to previous $6D$ reciprocal lattice treatments. The paper is organized as follows. In section \ref{sec2} we briefly review the generation of the $3D$ Penrose tiling which allows us to introduce the hard-sphere decoration. Section \ref{sec3} summarizes the generalized effective liquid approximation for the determination of the free energy of the quasicrystal. Our results are presented in section \ref{sec4} together with a discussion concerning the evaluation of quasilattice sums, while in the final section \ref{sec5} we gather our conclusions. \section{The icosahedral quasilattice} \label{sec2} In what follows, we consider the icosahedral quasilattice obtained by projecting a subset (to be specified below) of the $6D$ simple cubic lattice onto a $3D$ hyperplane\cite{Elser} $X_{{\scriptscriptstyle\parallel}}$. The orientation of $X_{{\scriptscriptstyle\parallel}}$ relative to the lattice is determined by requiring that the projected vectors $\{\mbox{{\bf e}}_{{\scriptscriptstyle\parallel}}^j \}$ of the $6D$ basis vectors $\{\mbox{{\bf e}}^j\}$ ($j=1,2,...6$) coincide with the six vertex axes of the icosahedron, i.e.: \begin{equation} \label{21} \mbox{{\bf e}}_{{\scriptscriptstyle\parallel}}^j=P_{{ \scriptscriptstyle \parallel} }^{jk}\mbox{{\bf e}}^k \end{equation} where a summation over repeated labels is understood. The matrix representation of the projection operator $P_{{\scriptscriptstyle \parallel}}$ is given by: \begin{equation} \label{22} P_{{\scriptscriptstyle\parallel}} =\frac{1}{\sqrt{20}} \left(\begin{array}{rrrrrr} \sqrt{5} & 1 & 1 & 1 & 1 & 1 \\ 1 & \sqrt{5} & 1 & -1 & -1 & 1 \\ 1 & 1 & \sqrt{5} & 1 & -1 & -1 \\ 1 & -1 & 1 & \sqrt{5} & 1 & -1 \\ 1 & -1 & -1 & 1 & \sqrt{5} & 1 \\ 1 & 1 & -1 & -1 & 1 & \sqrt{5} \\ \end{array}\right) \end{equation} and an elementary calculation leads to $|\mbox{{\bf e}}^j_ {\scriptscriptstyle\parallel}|^2= 1/2$ ($j=1,2,...6$) and \[ \cos({\bf e}^1_{\scriptscriptstyle\parallel},{\bf e}^j_ {\scriptscriptstyle\parallel})=\sqrt{5}/5\;\;\; (j=2,...,6) \] \[ \cos({\bf e}^{2+j}_{\scriptscriptstyle\parallel},{\bf e}^{2+k}_ {\scriptscriptstyle\parallel})=\left\{ \begin{array}{rll} \sqrt{5}/5 & \;\;\;\;\;\;\;\; & j-k=\pm1,\, \pm4 \\ - \sqrt{5}/5 & \;\;\;\;\;\;\;\; & j-k=\pm 2,\, \pm 3 \end{array} \right. \] showing that the vectors $\{\mbox{{\bf e}}_{{\scriptscriptstyle \parallel}}^j\}$ may be identified with the six vertex directions of the icosahedron. We also consider the $3D$ hyperplane perpendicular to $X_{{\scriptscriptstyle\parallel}}$, $X_{{\scriptscriptstyle\perp}}$, obtained upon projection of the $6D$ basis vectors $\{\mbox{{\bf e}}^j\}$ by the complementary projector $P_{{\scriptscriptstyle \perp}}$: \begin{equation} \label{23} \mbox{{\bf e}}^j_{{\scriptscriptstyle\perp}}= P_{{\scriptscriptstyle\perp}} ^{jk}\mbox{{\bf e}}^k; \,\,\,\,\,\,\,\, P_{{\scriptscriptstyle\perp}}^{jk} = \delta^{jk} -P_{{\scriptscriptstyle \parallel}}^{jk} \end{equation} where $\delta^{jk}$ is the Kronecker delta. It can be readily shown that the projected vectors $\mbox{{\bf e}}_{{\scriptscriptstyle\perp}}^j$ may also be identified with the six vertex directions of the icosahedron, but permuted with respect to the projected vectors $\mbox{{\bf e}}_{ {\scriptscriptstyle\parallel}}^j$, i.e. $\mbox{{\bf e}}_{{\scriptscriptstyle\perp}}^j\cdot\mbox{{\bf e}}_{{ \scriptscriptstyle\perp}}^k= -\mbox{{\bf e}}_{{\scriptscriptstyle\parallel}}^j\cdot\mbox{{\bf e}}_{ {\scriptscriptstyle\parallel}}^k$ ($j\neq k$). Both projections are dense in the $3D$ space but a quasilattice of finite density can be constructed by projecting onto $X_{{\scriptscriptstyle\parallel}}$ only those points whose perpendicular space projection lies within a bounded region $\chi$ known as the acceptance domain\cite{Elser}. To construct this bounded region we take the twenty distinct triplets $\{\mbox{{\bf e}}_{{\scriptscriptstyle\perp}}^i,\mbox{{\bf e}}_{{ \scriptscriptstyle\perp}}^j,\mbox{{\bf e}}_{{\scriptscriptstyle\perp}}^k\}$ of the projected vectors in $X_{{\scriptscriptstyle\perp}}$. Each triplet defines a rhombohedron of volume $v_{ijk}=|\mbox{{\bf e}}_{{\scriptscriptstyle\perp}}^i\times\mbox{{\bf e}}_{{ \scriptscriptstyle \perp}}^j\cdot\mbox{{\bf e}}_{{ \scriptscriptstyle\perp}}^k|$. It is easily found that half of these rhombohedra are ``large'', i.e. $v_{ijk}=\sqrt{8}\,\mbox{sin}(2\pi/5)/10$ and half ``small'', i.e. $v_{ijk}=\sqrt{8}\,\mbox{sin}(4\pi/5)/10$. The disjoint union of these twenty rhombohedra defines a closed convex region $\chi$ named the {\em triacontahedron} of volume $v=\sqrt{8}\,[\mbox{sin}(2\pi/5)+\mbox{sin}(4\pi/5)]$. The triacontahedron is therefore the projection onto $X_{{ \scriptscriptstyle\perp}}$ of the unit cell of the $6D$ cubic lattice. The selection of a subset of the $6D$ simple cubic lattice is accomplished by requiring that the orthogonal projections \{${\bf r}_{{ \scriptscriptstyle\perp}}^j\}$ of the lattice points \{${\bf r}^j\}$ of the $6D$ simple cubic lattice lie within the triacontahedron\cite{nota}. This construction yields a quasilattice in $X_{{\scriptscriptstyle\parallel}}$ which may also be regarded as a tiling of the $3D$ space by two kinds of rombohedra\cite{Henley}. Using a one-parameter Gaussian approximation for the density peaks of the quasilattice, the one-particle density can be written as: \begin{equation} \label{24} \rho({\bf r}) = \left(\frac{\alpha}{\pi}\right)^{3/2}\sum_j W({\bf r}_{{\scriptscriptstyle\perp}}^j) \mbox{e}^{-\alpha({\bf r} -{\bf r}_{ {\scriptscriptstyle\parallel}}^j)^2} \end{equation} where the sum runs over the Bravais lattice vectors of the $6D$ simple cubic crystal, $\alpha$ is the inverse width of the Gaussians and the weight function $W({\bf r}_{{\scriptscriptstyle \perp}}^j)$ is defined by: \begin{equation} \label{25} W({\bf r}_{{\scriptscriptstyle\perp}}^j)=\left\{\begin{array}{cl} 1\;\;\;\;\; & {\bf r}_{{\scriptscriptstyle\perp}}^j\in\chi\\ 0\;\;\;\;\; & \mbox{otherwise}\end{array}\right. \end{equation} It can be shown that the first three neighbour separations of the icosahedral quasilattice are\cite{Henley}: $r_1^2=(3-6\sqrt{5}/5)a^2$, $r_2=a$ and $r_3^2= (2-2\sqrt{5}/5)a^2$ where $a=\sqrt{2}/2$. Their average coordination numbers are $2/\tau^2$, 6 and 6, respectively, with $\tau$ denoting the golden ratio $\tau=(1+\sqrt{5})/2$. By locating a hard-sphere of diameter $\sigma=r_1$ at every vertex of the quasilattice, the packing fraction (the fraction of the total volume occupied by the spheres) is $\simeq 0.14$, i.e. a packing fraction characteristic of a fluid phase. On the other hand, if we look for accomodating hard spheres of diameter $\sigma=a$ at every vertex, the short distance $r_1$ does not allow this. But since the frequency of the nearest-neighbor separation is small, a better packing of hard spheres in the quasilattice can be obtained if one of the two vertex of each $r_1$-bond is left vacant. These short bonds form closed rings of 10 links and chains of even or odd links. If we also require that two adjacent sites cannot be empty there are only two ways of placing the hard spheres on rings and chains. In the present investigation we have randomly located a hard sphere on a vertex of every ring and in an endpoint of each chain. This determines the accomodation of the remaining hard spheres, the effect of taking different initial localizations having a negligible effect on our results as the number of vertex of the quasilattice increases. It can be shown\cite{Henley} that with this procedure, the volume occupied by the hard spheres increases leading to a hard-sphere packing fraction $\eta\simeq 0.629$, which is close to the random packing fraction ($\simeq 0.64$) and below the packing fraction of the crystal structures ($\simeq 0.74$ and $\simeq 0.68$ for the fcc and bcc crystals, respectively). This is, to our knowledge, the best icosahedral packing fraction of identical hard spheres reported so far. \section{Free energy of the hard-sphere icosahedral quasilattice} \label{sec3} In recent years, the freezing of hard spheres into perfect crystals has been successfully described by several nonperturbative density functional theories. We here consider the stability of the hard-sphere icosahedral quasilattice described in \ref{sec2} within one of such approaches, the generalized effective liquid approximation, which is now briefly summarized. The Helmholtz free energy $F$ of a solid characterized by a one-particle density $\rho({\bf r})$ is a functional of $\rho({\bf r})$, denoted by $F=F[\rho]$, which can be split as $F[\rho] =F_{\mbox{id}}[\rho]+F_{\mbox{ex}}[\rho]$ where \begin{equation} \label{31} \beta F_{\mbox{id}}[\rho] = \int\,d{\bf r}\rho({\bf r}) [\ln\{\Lambda^3\rho({\bf r})\}-1] \end{equation} is the ideal contribution with $\beta=1/k_BT$ the inverse temperature and $\Lambda$ the thermal de Broglie wavelength and \begin{equation} \label{32} \beta F_{\mbox{ex}}[\rho] = - \int\,d{\bf r}\rho({\bf r})\,\int\, d{\bf r}^{\prime} \rho({\bf r}^{\prime})\,\int_0^1\,d\lambda\,(1-\lambda)\, c({\bf r},{\bf r}^{\prime};[\lambda\rho]) \end{equation} is the excess term. In (\ref{32}) $c({\bf r},{\bf r}^{\prime};[\lambda\rho])$ is the direct correlation function of the solid and $\lambda$ ($0\leq\lambda\leq 1$) is a parameter defining a linear path of integration in the space of density functions $\rho_{\lambda}({\bf r})=\lambda\rho({\bf r})$ connecting a zero reference density to the one-particle density $\rho({\bf r})$ of the solid. The equilibrium solid density $\rho({\bf r})$, determined by functional differentiation, is the minimimum value of $F[\rho]$ at constant average density. This variational calculation implies the direct correlation function of the solid which is the only unknown in (\ref{31}-\ref{32}) and hence some explicit approximations for $F_{\mbox{ex}}[\rho]$ are required. Based on the similarity of the thermodynamic properties of the solid and fluid phases, the generalized effective liquid approximation first maps the excess free energy per particle of the solid onto that of some effective liquid, i.e.: \begin{equation} \label{33} \frac{1}{N} \int\,d{\bf r}\rho({\bf r})\,\int\,d{\bf r}^{\prime} \rho({\bf r}^{\prime})\,\int_0^1\,d\lambda\,(1-\lambda)\, c({\bf r},{\bf r}^{\prime};[\lambda\rho]) = \hat{\rho}\,\int\,d{\bf r}\,\int_0^1\,d\lambda\,(1-\lambda)\, c(|{\bf r}|;\lambda\hat{\rho}) \end{equation} where $\hat{\rho}$ is the density of the effective liquid, $N=\int\,d{\bf r}\rho({\bf r})$ is the number of particles, and $c(|{\bf r}|;\lambda\hat{\rho})$ is the direct correlation function of the liquid. Equation (\ref{33}) is referred to as the thermodynamic mapping. In a second step, the generalized effective liquid approximation defines an structural mapping in which the direct correlation function of the solid is mapped onto that of a liquid. However, this mapping cannot be done directly because the direct correlation function of the liquid is translationally invariant while that of the solid is not. But taking into account that in (\ref{32}) the direct correlation function of the solid appears doubly weighted by the solid density , the difficulty is overcome by defining the structural mapping as: \begin{equation} \label{34} \int\,d{\bf r}\rho({\bf r})\,\int\,d{\bf r}^{\prime} \rho({\bf r}^{\prime})\, c({\bf r},{\bf r}^{\prime};[\rho]) = \int\,d{\bf r}\rho({\bf r})\,\int\,d{\bf r}^{\prime} \rho({\bf r}^{\prime})\, c(|{\bf r}-{\bf r}^{\prime}|;\hat{\rho}[\rho]) \end{equation} With (\ref{32}-\ref{34}) a self-consistent non-linear integral equation is obtained for the determination of $\hat{\rho}[\rho]$ in terms of $\rho({\bf r})$ and the direct correlation function of the liquid. The complicated functional dependence $\hat{\rho}[\rho]$ can be simplified if $\rho({\bf r})$ is described in terms of a single order parameter $\alpha$ as in (\ref{24}) in which case $\hat{\rho}$ becomes an ordinary function of $\alpha$. The equilibrium solid density (i.e. $\alpha$) is then determined by minimizing at constant average density the solid free energy with respect to the Gaussian width parameter $\alpha$ for a given crystal structure. As explained elsewhere\cite{Tejero}, the non-linear integral equation for the determination of $\hat{\rho}[\rho]$ can be further transformed into a system of two coupled nonlinear differential equations in $\hat{\eta}(\lambda)$: \begin{equation} \label{35} \hat{\eta}^{\prime}(\lambda) = \frac{z(\lambda)-\psi(\hat{\eta}(\lambda))} {\lambda\psi^{\prime}(\hat{\eta}(\lambda))} \end{equation} and $z(\lambda)$: \begin{equation} \label{36} z^{\prime}(\lambda)=\Phi(\hat{\eta}(\lambda)) \end{equation} where $\hat{\eta}(\lambda) = \pi\hat{\rho}(\lambda)\sigma^3/6$ is the effective liquid packing fraction and $\sigma$ is the hard-sphere diameter. Using the one-parameter approximation (\ref{24}) for the one-particle density of the quasilattice, $\Phi(\hat{\eta}(\lambda))$ is given by: \begin{equation} \label{37} \Phi(\hat{\eta}(\lambda))= -\frac{1}{N} \sum_{i=1}^{N}\sum_{j=1}^{N}\,W({\bf r}_{{\scriptscriptstyle \perp}}^i)\,W({\bf r}_{{\scriptscriptstyle\perp}}^j) \int_0^{\infty}\, dR\,R\, c(R;\hat{\eta}(\lambda)) S(R;\alpha,r_{ij}) \end{equation} where $r_{ij}= |{\bf r}_{{\scriptscriptstyle\parallel}}^i-{\bf r}_{ {\scriptscriptstyle\parallel}}^j|$ and \begin{equation} \label{38} S(R;\alpha,r_{ij}) =\left[\frac{\alpha}{2\pi r_{ij}^2}\right]^{1/2} \left[ \exp(-\alpha(R-r_{ij})^2/2)-\exp(-\alpha(R+r_{ij})^2/2)\right] \end{equation} In (\ref{35}) and (\ref{36}) the prime denotes the derivative with respect to the argument and $\psi(\hat{\rho})/\beta$ is the excess free energy per particle of the fluid phase. For the latter we will use the Carnahan-Starling compressibility factor to obtain $\psi$ by thermodynamic integration of the equation of state while the Percus-Yevick equation is used for the structure of the fluid phase, i.e. the direct correlation function. Eqs. (\ref{35}) and (\ref{36}) have to be integrated numerically from $\lambda=0$ to $\lambda=1$ with initial conditions $\hat{\eta}(0) =z(0)=0$ and the excess free energy per particle of the quasilattice (\ref{32}) is finally determined as $\psi(\hat{\eta}(1))$. \section{Results} \label{sec4} Before looking for a numerical solution of (\ref{35}-\ref{36}), we deal with a delicate point concerning the convergence of the quasilattice sums in (\ref{37}). In order to emphasize it let us rewrite the rigth hand side of eq. (\ref{37}) in the form: \begin{equation} \label{41} \frac{1}{N}\sum_{i=1}^{N}\sum_{j=1}^{N}A(r_{ij}) \end{equation} We first note that for a Bravais lattice (\ref{41}) reduces to, \begin{equation} \label{42} \sum_{j}\, z_j A(r_j) \end{equation} where the sum runs over spherical shells of sites centered around the site at the origin, $r_j$ is the distance of shell $j$ to the origin, and $z_j$ is the number of sites at the $jth$-shell. In passing from (\ref{41}) to (\ref{42}) the translational symmetry of the Bravais lattice has been used. Moreover, since $A(r_j)$ decreases rapidly with distance, only a relatively small number of shells around the origin give a nonnegligible contribution to the sum leading to a rapid convergence of (\ref{42}). But if the lattice does not have translational symmetry, the evaluation of (\ref{41}) becomes a delicate numerical problem if one looks for achieving a rapid convergence. For instance, if we consider a quasilattice of $N$ sites the convergence of (\ref{41}) (resulting from considering the $N(N-1)/2$ different pairs $i\neq j$) becomes so slow as $N$ increases that it remains unreachable through usual computational efforts. This is also the case for a Bravais lattice as it may be tested by evaluating the sum (\ref{41}) which, on the other hand, can be easily determined through (\ref{42}). As stated above, a possible way for dealing with the quasilattice sum\cite{Ash} is to use the $6D$ reciprocal lattice. However, this procedure only provides a partial solution to the convergence of the sum since it is necessary to truncate the sum at some maximum value of the reciprocal lattice vector leading to an estimate error in the free energies of about $2\%$. We here propose an alternative solution to the convergence problem of (\ref{41}) which provides a substantial reduction of errors and computation time in the determination of the quasicrystal free energy. By starting with the $6D$ simple cubic lattice, we construct the quasilattice using the projection formalism described in \ref{sec2}. Let $M>>1$ be the number of the lattice points generated in the quasilattice and draw an spherical surface containing almost all the lattice points. Inside the sphere we construct a concentric sphere with $N>>1$ ($N<M$) lattice points. By rewriting (\ref{41}) as: \begin{equation} \label{43} A(0) + \frac{1}{N}\sum_{i=1}^{N}\left[\sum_{j\neq i}A(r_{ij})\right] \end{equation} where we have separated the $r_{ij}=0$ contributions, we calculate the the $i\neq j$-terms by first choosing a lattice point $i$ ($i = 1,2,... N$) inside the sphere and then summing consecutively over all neighbouring $j$ lattice points until the relative error of the sum in brackets in (\ref{43}) is less than a prefixed value. For lattice points $i$ well inside the sphere, our procedure takes into account all the relevant ``interactions'' in (\ref{41}). However, if the lattice point $i$ lies near the surface of the sphere, our procedure overestimates the ``interactions'' in (\ref{41}) because the sum in brackets contains points outside the sphere. We have found that these boundary errors can be reduced by increasing $N$ (and therefore $M$), in such a way that for $M = 22000$ and $N = 10000$, the estimate error of the free energy per particle of the quasilattice is about $0.1\%$. Under such conditions, the computation time needed for evaluating the variational free energy (for each pair $\alpha$-$\eta$) is around $20' $ c.p.u. in a VAX 9000. In Fig. 1, the variational free energy per particle of the quasicrystal $\beta\phi=\beta F/N-3\ln(\Lambda/\sigma)+1$ is represented versus $Y=(\alpha\sigma^2/2)^{1/2}$ for different packing fractions. We have found minima for the quasicrystal free energy as a function of the Gaussian width parameter $\alpha$ for $\eta\geq 0.51$, i.e. a stable or metastable quasicrystalline phase. Similarly to ref.\cite{Ash} we have found that the quasicrystal turns out to be the more localized phase since the minima are always situated at greater $Y$-values than the compact fcc crystalline phase. Our results are gathered in Table \ref{tab1} where the free energy and the Gaussian width parameter at the free energy mimimum of the quasicrystal are compared to those of the fcc crystal. In Fig. 2 we represent the solid free energy per particle versus the packing fraction $\eta$ for the three solid phases (fcc, bcc and quasicrystal). We also include in the figure the fluid free energy obtained by thermodynamic integration of the Carnahan-Starling equation of state. It is seen that the quasicrystal is always metastable with respect to the remaining phases, the gaps of the free energy being somewhat less than those reported by M$^{c}$Carley and Ashcroft\cite{Ash}. \section{Conclusions} \label{sec5} We have analyzed the stability of a hard-sphere quasicrystal obtained from a simple decoration of the $3D$ Penrose tiling which has been designed for optimizing the packing fraction. The stability has been analyzed using the generalized effective liquid approximation. A simple method for evaluating the quasilattice sums in the $3D$ real space has been formulated. The method minimizes the boundary effects of finite quasilattices leading to a substantially better convergence than previous works. Our results show that the quasicrystal is metastable with respect to the crystalline and fluid phases. Such results agree with recent reported\cite{Ash} calculations for a hard-sphere quasicrystal obtained from the modified weighted density approximation. Therefore, within these nonperturbative density functional theories entropy is insufficient to stabilize one-component quasicrystals. Since all known quasicrystals have complex metallic alloy phase structures it has been argued that for the stability of quasicrystals it is neccesary to have at least two class of atoms. The generalization of the one-component quasicrystal structure to an ordered two-component structure has been investigated by M$^{c}$Carley and Ashcroft \cite{Ash} who concluded that small changes in the diameter ratio of the two-component hard-sphere quasicrystal are not a stabilizing factor. It should be expected that energetic contributions resulting from considering more realistic interactions would propitiate stability. This possibility seems to be ruled out using the well-known perturbation schemes when applied to quasicrystals in view of the great free energy differences between the crystalline and quasicrystalline hard-sphere phases. Thus, the stability of quasicrystal structures within the modern density functional approaches is at present an open question. \section{Acknowledgments} We thank R. Brito and J. M. R. Parrondo for useful discussions. H. M. Cataldo has been supported by grants PID 97/93 from Consejo Nacional de Investigaciones Cient\'{\i}ficas y T\'ecnicas (Argentina) and EX100 from Universidad de Buenos Aires. C. F. Tejero acknowledges the DGICYT, Spain (PB91-0378) for its financial support.
\section{MATERIALS AND METHODS} Wet-spun oriented samples were prepared from calf-thymus DNA (Pharmacia) with a molecular weight of $\sim 1.6 \times 10^7$ by the standard method \cite{bib12}. This spinning allows controlled production of sufficient amounts of highly oriented thin films by spooling DNA fibres which are continuously stretched during precipitation into an aqueous alcohol solution. Films of thickness of $\sim 0.5~\rm mm$ and surface area between $\rm 5$ and $10~\rm mm^2$ were used in the experiments reported here. Unoriented fibers of high MW ($\sim 1 \times 10^8$) DNA were prepared from whole adult chicken blood (Truslow Farms, Chestertown, MD) as described in McGhee {\sl et al.} \cite{bib11}. This DNA was further purified with three extractions against phenol/chloroform (50:50) and once with chloroform alone. Then DNA was ethanol precipitated in sodium acetate, pelleted by centrifugation, washed twice with 70\% ethanol and dried. This DNA was used in all preparations involving unoriented fibers. Oriented as well as unoriented DNA fibers were equilibrated with various solutions of PEG (20,000 MW) in 0.5 M NaCl, 10 mM Tris/ 1 mM EDTA, pH 7 in vast excess. Under these conditions, PEG (20,000 MW) is completely excluded from the DNA phase for concentrations greater than $\approx 7 \% $(w/w). The equilibration time was usually from four days to a week. Measurements on both orientationally ordered (wet-spun) as well as ``powder'' samples show that there is essentially no difference in osmotic pressure vs. concentration (interhelical spacing) dependence between the two preparations. The two preparations differ only in the size of the oriented domains. X-ray diffraction was performed at $20^{\circ}$C with an Enraf-Nonius Service Corp. (Bohemia, NY) fixed-anode FR 590 x-ray generator equipped with image plate detectors. Image plates were read and digitized by a Phosphor Imager (Molecular Dynamics, CA) and processed with NIH Image 1.55 program (W. Rasband, NIH, Bethesda, MD) modified by us. The position of the first order diffraction peaks ($r_{1.max}$) is obtained by radially averaging the scattering profile around the direct beam. Angular intensity profiles were taken at the position of the maximum of the first order diffraction peak and were then Fourier transformed to extract the bond orientational order parameter ${\cal C}_6$, {\sl i.e.} the sixth order Fourier coefficient. If there were perfect alignment of the x-ray beam and the average director of the oriented DNA sample , the angular dependence of the six-fold symmetric scattering function could be Fourier analyzed in terms of \cite{strandburg} \begin{equation} {\cal S}(\theta,r_{1.max}) = I_0(r_{1.max})\left[{\textstyle\frac{1}{2}} + \sum_{n=1}^{\infty} {\cal C}_{6n}~\cos{6n(\theta - \theta_0)}\right] + I_{BG}, \label{equ0} \end{equation} where $I_{BG}$ is the background intensity. Because the orientation was only approximate there was usually a small ${\cal C}_2$ component present in the Fourier analyzed angular profiles. We have rescaled the value of ${\cal C}_{6}$ to correct for this . DNA samples at various densities were sealed between microscope cover glass and were observed under a microscope (Olympus) equipped with crossed polarizers. The image was digitized and analyzed with NIH Image 1.55. The ``fingerprint'' cholesteric pattern \cite{biophys} with long fragment DNA was never as regular as is typical of short fragment DNA. Rather long DNAs achieve oriented domains of much smaller size. The $\rm ^{31}P$ NMR measurements were performed on a Bruker MSL-300 spectrometer (Billerica, MA) using a high power probe with a 5 mm solenoidal sample coil which was doubly tuned for $\rm ^{31}P$ (121.513 MHz) and protons (300.13 MHz). Gated broadband decoupled $\rm ^{31}P$ spectra were observed with a phase cycled Hahn echo sequence. A delay time between the 90 degree pulse and 180 degree pulse of 30 microseconds was chosen. Typically 20,000 to 80,000 scans with a recycle delay time of 1s were accumulated. Exponential linebroadening with a linewidth of 200 Hz was used. First moments of the NMR spectra ($M_1$) were calculated in standard fashion according to \begin{equation} M_1 = \frac{\int_{-\infty}^{+\infty} f(\omega)~\omega d\omega}{\int_{-\infty}^{+\infty} f(\omega)~d\omega}, \label{equM1} \end{equation} where $f(\omega)$ is the spectral intensity at the frequency $\omega$. The frequency of the center of the spectrum, determined as half height of the integral $\int_{-\infty}^{+\infty} f(\omega)~d\omega$, was set to zero. The measured dependence of the osmotic pressure of the DNA phase on DNA concentration allows one to evaluate the reversible work done at constant temperature, pressure and chemical potential of salt as the system is brought from an initial (i) to a final (f) configuration. The difference in free energy is \begin{eqnarray} \Delta {\cal G} &=& - \int^{V^{f}}_{V^{i}} \Pi(V_{DNA})~dV_{DNA}. \label{equ1} \end{eqnarray} The excess or packing energy per unit length of the DNA helix can now be obtained as \begin{equation} \frac{\Delta{\cal G}}{L} = - \sqrt{3} \int^{D_f}_{D_i}~\Pi(D)~DdD, \label{equ4} \end{equation} where $D$ is the interhelical spacing assuming the DNA array is at least locally hexagonal. Since the DNA osmotic pressure decays exponentially at small and intermediate values of $D$, a finite density interval is sufficient to evaluate the above integral to satisfactory accuracy. We have taken $D_i$ corresponding to the concentration $15 {\rm mg/ml}$ (data not shown on Fig.1), which marks the onset of the condensed (anisotropic) DNA phase \cite{kunal}. Since thermal fluctuations are contributing to the free energy it is reasonable to express the calculated free energy per unit length, $\frac{\Delta{\cal G}(D)}{L}$, in its ``natural'' units of $kT$ per persistence length ${\cal L}_p$ ( $\approx 500$ \AA). In these units one can write \begin{equation} \frac{\left(\Delta {\cal G}(D)/kT\right)}{{L}/{\cal L}_p} = \frac{{\cal L}_p}{\zeta (D)}, \label{extra1} \end{equation} where ${\zeta (D)}$ is the contour length of DNA associated with kT of packing energy in the condensed phase. \vfill \eject \section{RESULTS} \subsection{\sl Osmotic Stress Measurements} The dependence of osmotic pressure on the concentration of the unoriented DNA subphase has been investigated in detail \cite{bib9, bib10, bib14}. The corresponding interhelical spacings were obtained by measuring the first order x-ray diffraction peak on unoriented DNA samples assuming local hexagonal packing symmetry. This assumption was verified experimentally in the high density region (I) (see Fig.1) through the existence of weak higher order reflections and now by observing well developed six-fold symmetric bond orientational order (see section 2.2). Similar measurements were performed on oriented samples that show the same interaxial spacing (or density) dependence on $\Pi$ as the unoriented samples (see Fig. 1) and thus have the same free energy, within experimental error. There are two distinct regions in the ${\Pi} - D$ curve. In the high pressure regime, the interhelical distance does not depend on the salt concentration . The forces between helices in this region were interpreted as resulting from water - mediated structural forces \cite{bib9}. At lower pressures a sensitivity of D to salt concentration is clearly discernible. The effective decay length for the interhelical interactions, however, is about twice the predicted Debye screening length \cite{bib15} for salt concentrations $<$ 1.0 M, where electrostatic interactions are not overwhelmed by hydration forces. The two scaling regimes of the osmotic pressure are separated by a narrow crossover region in the ${\Pi} - D$ curve at about $ 32 - 34$~\AA. \subsection{\sl Packing Symmetry} The two regimes in the osmotic pressure curve are also clearly evident in the qualitative characteristics of the X-ray diffraction on oriented samples (see Fig. 2). For oriented samples of DNA in the high osmotic pressure regime (I) the cross section of the first order interaxial diffraction peak with the DNA helical axis oriented parallel to the incoming beam is a circular ring with six-fold modulation in the intensity which clearly reflects the six-fold symmetric long range bond orientational order of the underlying DNA lattice Fig.2 (inset). Azimuthal modulation of the first order diffraction peak at close DNA spacings has been observed previously in neutron diffraction studies \cite{bib13} with fibers of NaDNA and LiDNA at low excess salt content. As the osmotic pressure is lowered the six-fold modulation of the first order diffraction peak disappears and is unobservable below the transition, 32-34 \AA, region (see inset Fig.2) in the $\Pi - D$ curve. For spacings less than 35 ~\AA~ the changes in the six-fold modulation of the diffraction peak were reversible. However, once the bond orientational order is lost, it cannot be regained by simply increasing the osmotic pressure. The subsequent chain entanglement due to the looser nature of the packing in this low pressure phase apparently precludes the reestablishment of long range bond orientational order. The details of the first order diffraction peak are irretrievably lost leading to a circular powder pattern. The details of the azimuthal profile of the diffraction pattern were independent of the X-ray beam size up to cross sectional areas on the order of $\sim \rm mm^2$. The bond orientational order thus appears to be of very long range indeed. The translational order, on the other hand, estimated crudely from the radial linewidth of the first order diffraction peak \cite{chaikin} and extremely weak higher order reflections (J. R\" adler, personal communication), appears to be of a much shorter range, on the order of several lattice spacings. To quantify this change in orientational bond order, we have measured the azimuthal intensity distribution of the first order diffraction peak and extracted the corresponding Fourier coefficients shown in Fig.2. Generally the Fourier spectra showed pronounced peaks for ${\cal C}_n$ with $n = 0 ~{\rm and}~ 6$, with typically a small, but discernible additional contribution from ${\cal C}_2$, most probably reflecting a slight misorientation of the x-ray beam direction and the average director of the oriented DNA sample. The extracted ${\cal C}_6$ coefficients, that are also corrected for misalignment, shows a gradual loss of lateral bond orientational order as the DNA density passes from the high to low osmotic pressure regimes . The nature of the low osmotic pressure phase can be further ascertained by polarized light microscopy which clearly reveals the existence of a ``fingerprint'' texture characteristic of a cholesteric phase \cite{Livolant}. Though the pitch of the cholesteric phase varies with density of the DNA phase in the vicinity of I $\longrightarrow$ II transition, we were unable to quantify this accurately because the orientational domain sizes were, in general, small. Due to the high molecular weight of the DNA, the samples could not be manipulated by an applied external orienting magnetic field to increase the domain size. \subsection{\sl Phosphate Backbone Dynamics} An earlier analysis of the $\Pi - D$ curve suggested that there was a relatively sudden change in lattice fluctuations, inferred from changes in x-ray scattering peak widths, within the 32-34 \AA\ transition region, \cite{bib10}. This change in motion can now be seen very clearly in the $\rm ^{31}P$ NMR spectra. The insert in Fig. 3 shows two $\rm ^{31}P$ NMR spectra - one within the high pressure regime and one in the low pressure, cholesteric phase - that clearly demonstrate a symmetry change in the effective tensor of chemical shift. While any quantitative relation between the values of the effective tensor of chemical shift, the spectral first moment, and the details of the molecular motions is highly model dependent, it is clear that there is a qualitative difference in the DNA dynamics between the two pressure regimes. We have quantified this change by analyzing the first moments of the $\rm ^{31}P$ NMR spectra, shown in Fig.3. If there are no other processes contributing to resonance broadening and since osmotically equilibrated samples are monophasic, the observed increase of the first moment as the interaxial spacing decreases is due to a decreased mobility of DNA helices. The molecules are obviously immobilized to a substantial degree in the high pressure phase though the spectral first moment does decrease somewhat , see Fig.3, as the 32-34 \AA\ transition region is approached. In the low pressure region, the phosphate mobility appears to be significantly greater. The difference in the spectral first moments between the two pressure regimes suggests a drastic increase of motional amplitudes for the cholesteric phase but the mobility appears not to change substantially with density within this phase. Typical principal values of the chemical shift tensor extracted from spectra at high applied osmotic stress are $\sigma_{xx} \approx -60$ ppm, $\sigma_{yy} \approx -5$ ppm, and $\sigma_{zz} \approx 65$ ppm. Comparable values for essentially completely immobilized dry DNA, are $-83$ ppm, $-22$ ppm and $110$ ppm (measured relative to 85\%~ phosphoric acid as a standard) \cite{shindo}. The effective tensor for DNA in the high pressure regime shows that phosphate motions are quite restricted. No fast rotation around one axis is present because this would have resulted in a tensor with axial symmetry. \vfill \eject \section{DISCUSSION} \subsection{\sl Structure and Dynamics} This study, together with earlier measurements of intermolecular forces \cite{bib9,bib10}, presents a departure from the usual gravimetric method of sample preparation. By bringing ordered phases into equilibrium with large "reservoirs" of salt-plus-polymer solutions rather than by making stoichiometric mixtures of salt, water and DNA, it is possible to set all the intensive thermodynamic variables associated with the resulting single liquid-crystalline phase. These simultaneous measurements of the structure, motion and thermodynamic functions of DNA phases have focused on high density DNA phases (with interhelical spacings between about 25 and 55 \AA) at one ionic strength (0.5 M NaCl). This density region extends from $\sim 120~\rm mg/ml$ to $\sim 600~\frac{mg}{ml}$. At lower densities there is a transition to a cholesteric phase from one of the (presumably) blue phases \cite{amelie}, while at higher densities there is a transition into a three dimensional crystal with a simultaneous B $\longrightarrow$ A transition in DNA conformation. For the long fragment DNA investigated here, the isotropic $\longrightarrow$ anisotropic transition is still quite remote ($\sim 10~\rm mg/ml$ \cite{kunal}). The structural, dynamic, and osmotic stress data presented here are all consistent with the existence of two different DNA phases separated by a transition region at a DNA density of $\sim 320 - 360 {}~\rm mg/ml$, corresponding to interhelical spacings of $\sim 32 - 34$~\AA. Previous work on short fragment ($146~{\rm bp} \sim 500$~\AA~ long) DNA \cite{bib3} also gave clear evidence for the existence of a series of structurally distinct regions as a function of DNA concentration. The transition from a cholesteric to a 2D-hexagonal phase for short fragment DNA was observed at $\sim 32$~\AA. Remarkably the $\sim 32$~\AA~ interaxial spacing is also close to the spacing from which ${\rm Mn}^{2+}$ or ${\rm Co}^{3+}$-DNA collapses in a first order transition under osmotic stress \cite{don}. Is this a distance at which the details of the chiral double-helical structure come to be sensed in molecular interaction? What these experiments do not show clearly is the nature and the order of the transition between bond orientationally ordered and cholesteric phases. There is no detectable discontinuity ( the accuracy of the measurement of the interhelical spacing in this regime of DNA densities is $\sim 1$ \AA ) in the $\Pi$ {\sl vs.} interaxial separation curve that is seen when DNA makes a clear first-order transition \cite{don}. This should not be taken as definitive evidence, however, that the transition is second order. An extremely narrow phase coexistence window could simply be a property of polymers in liquid crystalline mesophases \cite{edwards}. The accuracy of the azimuthal scans of the first order diffraction peak as well as the first moment of the $\rm ^{31}P$ NMR spectra also precludes a definitive measure of the order of the transition. \subsection{\sl Free Energy and Intermolecular Forces} The "osmotic stress" exerted by the excluded polymer is the rate of change of free energy with change in the amount of solution in the DNA phase, i.e., $\Pi = -{\partial G}/{\partial V_{DNA}}$. By integrating the osmotic pressure curve one thus obtains the change in the system free energy, Eq.\ref{equ4}. In the insert to Fig. 1 we have plotted this free energy as a function of molecular separation. It is given in thermal units of kT per persistence length (see Eq.\ref{extra1}), and spans a wide range of energy scales, from about kT per 2.5~\AA~ at $\log(\Pi) \sim 8~{\rm dynes/cm^2}$ to about kT per 100~\AA~ at $\log(\Pi) \sim 6~{\rm dynes/cm^2}$. Previous work \cite{bib14} has established that forces in the high pressure regime are dominated by exponentially decaying hydration interactions with a decay length $\lambda \sim 3 - 4$~\AA~ that is basically independent of the ionic strength. In the low pressure regime, the interaxial spacing dependence on osmotic stress is also exponential, but the effective decay length is about twice the expected Debye decay length (at least for salt concentrations between {}~ 0.2 and $\sim 0.8$~ M). The enhanced decay length and a rescaling of the strength of the interactions between DNA helices in this regime of DNA densities was shown to be due to the progressive onset of conformational disorder characterized by the fluctuations in the mean position of the molecules along the average director \cite{bib14, flu} and deduced from the width of the interhelical x-ray scattering peaks. The switch between fluctuation enhanced forces and bare potentials was not found to be gradual, but rather quite abrupt as the DNA density passed the $\sim 340~\rm mg/ml$ limit \cite{bib10}, correlating nicely with emergence of longitudinal order between helices seen in the studies of Livolant {\sl et al.} \cite{bib3} on short fragment DNA, as well as with the onset of lateral bond orientational order and broadening of the phosphate NMR peak reported here. The fluctuation-enhanced effective interactions observed in DNA arrays have the same origin as the effective interactions in smectic arrays. They are due to the interplay between conformational fluctuations and bare short range potentials \cite{bib15}. The clearly emerging enhancement of electrostatic decay length to about twice the Debye length, not yet so easily seen in lipid bilayer smectic arrays, could be connected with the different dimensionalities of the two systems (2D periodicity {\sl vs.} 1D periodicity). \subsection{\sl Perspectives and Directions} Molecular interactions in DNA arrays, extracted from the measured osmotic pressure of the array, are expected \cite{bib10} to vary with the interaxial spacing $D$ as $\sim K_0(D/\lambda)$ , where $ K_0(x)$ is the modified Bessel function with asymptotic behavior $ K_0(x) \approx (\pi /2x)^{1/2}~e^{-x}$, with a decay length $\lambda$ dependent on the salt concentration \cite{bib14}. In this respect, as noted by Nelson \cite{nelson2}, the interactions between helices in condensed DNA mesophases are formally and surprisingly closely related to the interactions between magnetic vortex lines in flux-line lattices of high-$\rm T_c$ superconductors which, apart from the lack of hard core repulsions, share the same form of interaction potential. The existence of a line (polymer) hexatic phase, intermediate between a line crystal and a line liquid, was hypothesized by Marchetti and Nelson \cite{nelson1} specifically for the case of magnetic flux-line lattices. It appears that the bond ordered DNA phase (region I) described above is perhaps this type of intermediate phase. The transition from a line hexatic phase in DNA into one of the possible less ordered phases is complicated by the presence of chiral coupling in the molecular interactions at lower densities, leading to the cholesteric, not a line liquid, phase. The occurrence of line hexatic between the cholesteric and the crystalline (A-form DNA) phases makes it difficult to compare directly with existing theoretical predictions. Its existence nevertheless introduces a new possible scenario into the melting sequence of ordered polyelectrolyte arrays. To say that DNA provides an opportunity to learn about liquid-crystals is not to say that it has already given clear answers to basic questions. What is the nature of the transition from a phase with well developed six-fold symmetric bond orientational order to a a skewed, cholesteric phase when the molecules are allowed to move apart? Why does this change in symmetry couple with the molecular motions that cause extra interaxial separation \cite{bib10}? What is the nature of molecular packing in the long polymer cholesteric phase compared to the more common twisted nematic phases of shorter molecules? These combined structural studies \cite{bib3}, osmotic stress measurements of free energies, and x-ray \& NMR probes of molecular disorder and motion now provide a direction and an opportunity for further development of systematic theoretical analyses. \vfill \eject
\chapter{Introduction} Cosmology has over the past fifteen years emerged as a vibrant and exciting subfield of physics. It is based on the marriage of quantum field theory and particle physics on the one hand with classical general relativity on the other. One of the main goals of modern cosmology is to explain the structure of the Universe on the scale of galaxies and beyond. Thus, the experimental/observational basis of the field lies in astronomy, and there is a lot of interaction between theoretical cosmologists and observational astronomers and astrophysicists. The main goal of these lectures is to give an introduction to the two most developed classes of structure formation theories: those based on inflation and those based on topological defects. I will give a brief survey of relevant observational results from large-scale structure surveys and from searches for cosmic microwave background (CMB) anisotropies, and I will attempt a preliminary comparison with theoretical predictions. A summary of some recent speculative ideas concerning the connection possible solutions of the singularity problem of classical cosmology will be presented. These notes are intended as a pedagogical introduction rather than as a comprehensive review. For comprehensive discussions of inflation, the reader is referred to Refs. 1-3, and for detailed reviews of topological defect models to Refs. 4-7. An introduction to quantum field theory methods used in modern cosmology can be found in Ref. 8. These notes are an updated and expanded version of earlier lecture notes$^{9)}$ and draw on material presented elsewhere$^{10)}$ in which some of the topics are treated in more detail. I hope to persuade the reader that cosmology is an exciting area of physics with close connections to particle and high energy physics, and with a steady stream of new data from astronomy and astrophysics. There is also a close connection with fundamental physics. In fact, cosmology may well be the only arena in which theories such as superstring theory are testable. The outline of these lectures is as follows: Section 2 is a review of standard cosmology, focusing on its basic principless, its observational support and its problems. Section 3 is a brief overview of ``new cosmology." I argue why, to obtain an improved cosmological scenario, we need to treat matter using particle physics and field theory. Next, I introduce the basic idea of the inflationary Universe scenario and explain how it leads to a solution of some of the problems of standard cosmology. In particular, it provides a mechanism for the formation of structure in the Universe. The section continues with a brief introduction to the topological defect models of structure formation, an explanation for the need of dark matter, and a survey of the present models. In Section 4, I present the basics of structure formation, beginning with a survey of some of the relevant large-scale structure data. Structure in the Universe is assumed to grow by gravitational instability. I summarize the essentials of the Newtonian theory of cosmological perturbations (valid on length scales smaller than the apparent horizon (Hubble radius)) and of the relativistic theory$^{11)}$ (required to study scales beyond the horizon). I also discuss free streaming. Section 5 contains an overview of inflationary Universe models and of the mechanism for the generation and evolution of perturbations which they provide. Section 6 presents an overview of topological defect models of structure formation. To begin, a classification of defects is given. Next it is shown that in models which admit topological defects, they are inevitably formed during a symmetry breaking phase transition$^{12)}$. Cosmic string and global texture models of structure formation are discussed in detail. In particular, it is pointed out that if defects are responsible for seeding galaxies, there must be new physics at a scale of $\eta \sim 10^{16}$ GeV. Section 7 focuses on CMB anisotropies. It is shown why theories of structure formation inevitably produce such anisotropies, the predictions of the various models are reviewed, and a comparison with recent observations is given. Finally, Section 8 contains a summary of a modified theory of gravity in which many of the singularities of classical cosmology can be smoothed out. The theory is based on a {\it limiting curvature construction}. Also discussed are some possible connections between superstring theory and cosmology. In particular, a mechanism which might single out three large spatial dimensions is suggested. In this writeup, units in which $c = \hbar = k_B = 1$ are used unless mentioned otherwise. The space-time metric $g_{\mu\nu}$ is taken to have signature $(+,-,-,-)$. Greek indices run over space and time, latin ones over spatial indices only. The Hubble expansion rate is $H (t) = \dot a (t) / a (t)$, with $a(t)$ the scale factor of a Friedmann-Robertson-Walker (FRW) Universe. The present value of $H$ is $100 h$ kms$^{-1}$ Mpc$^{-1}$, where $0.4 < h <1$. Unless stated otherwise, the value of $h$ is taken to be 0.5. The cosmological redshift at time $t$ is denoted by $z(t)$. As usual, the symbols $G$ and $m_{pl}$ stand for Newton's constant and Planck mass, respectively. Distances are measured in pc (``parsec") or Mpc, where 1 pc corresponds to 3.1 light years. \chapter{Review of Standard Cosmology} \section{Principles} The standard big bang cosmology rests on three theoretical pillars: the cosmological principle, Einstein's general theory of relativity and a perfect fluid description of matter. The cosmological principle$^{13)}$ states that on large distance scales the Universe is homogeneous. From an observational point of view this is an extremely nontrivial statement. On small scales the Universe looks rather inhomogeneous. The inhomogeneities of the solar system are obvious to everyone, and even by the naked eye it is apparent that stars are not randomly distributed. They are bound into galaxies, dynamical entities whose visible radius is about $10^4$ pc. Telescopic observations show that galaxies are not randomly distributed, either. Dense clumps of galaxies can be identified as Abell clusters. In turn, Abell cluster positions are correlated to produce the large-scale structure dominated by sheets (or filaments), with typical scale 100 Mpc, observed in recent redshift surveys$^{14)}$. Until recently, every new survey probing the Universe to greater depth revealed new structures on the scale of the sample volume. In terms of the visible distribution of matter there was no evidence for large-scale homogeneity. This situation changed in 1992 with the announcement$^{15)}$ that a new redshift survey, complete to a depth of about $500 h^{-1}$ Mpc, had discovered no prominent structures on scales larger than $100 h^{-1}$ Mpc. This is the first observational evidence from optical measurements in favor of the cosmological principle. However, to put this result in perspective we must keep in mind that the observed isotropy of the CMB temperature$^{16)}$ to better than $10^{-5}$ on large angular scales has been excellent evidence for the validity of the cosmological principle. The second theoretical pillar is general relativity, the theory which determines the dynamics of the Universe. According to the cosmological principle, space at any time $t$ is a three dimensional surface with maximal symmetry (translations and rotations). There are three families of such spaces$^{17)}$: flat Euclidean space $R^3$, the three sphere $S^3$, and the hypersphere $H^3$. The proper distance $ds^2$ on these three surfaces can be written in spherical coordinates as $$ ds^2 = a(t)^2 \, \left[ {dr^2\over{1-kr^2}} + r^2 (d \vartheta^2 + \sin^2 \vartheta d\varphi^2) \right] \, . \eqno\eq $$ The constant $k$ is $+1, \, 0$, or $-1$ respectively for $S^3, \, R^3$ and $H^3$. The Einstein equations of general relativity imply that $a(t)$ -- called the scale factor of the Universe -- evolves in time. The proper distance/time in space-time is $$ ds^2 = dt^2 - a(t)^2 \left[ {dr^2\over{1-kr^2}} + r^2 (d \vartheta^2 + \sin^2 \vartheta d\varphi^2) \right] \, . \eqno\eq $$ By a coordinate transformation, $a(t)$ can be set equal to 1 at the present time $t_0$. \smallskip \epsfxsize=8cm \epsfbox{bfig1.eps} {\baselineskip=13pt \noindent{\bf Figure 1:} Sketch of the expanding Universe. Concentric circles indicate space at fixed time, with time increasing as the radius gets larger. Points at rest have constant comoving coordinates. Their world lines are straight lines through the origin (e.g. $L$).} \medskip To obtain a simple visualization of an expanding Universe, consider space to be the surface of a balloon. We draw a grid on the surface and use it to define coordinates $\underline{x}^c$ (the superscript $c$ stands for comoving). Points at rest on the surface of the balloon have constant comoving coordinates. However, if the balloon is being inflated, then the physical distance $\Delta x^p$ betwen two points at rest with comoving separation $\Delta x^c$ increases: $$ \Delta x^p = a(t) \Delta x^c \, . \eqno\eq $$ The scale factor $a(t)$ is proportional to the radius of the balloon (see Fig. 1). According to Einstein's equivalence principle, particles in the absence of external nongravitational forces move on geodesies, curves which extremize $ds^2$. The velocity of a particle relative to the expansion of the Universe is called peculiar velocity $v_p$ $$ v_p = a(t) \, {dx^c\over dt} \eqno\eq $$ and obeys the equation $$ \ddot v_p + {\dot a\over a} \, v_p = 0 \, , \eqno\eq $$ from which it follows that $$ v_p (t) \sim a^{-1} (t) \, . \eqno\eq $$ The dynamics of an expanding Universe is determined by the Einstein equations, which relate the expansion rate to the matter content, specifically to the energy density $\rho$ and pressure $p$. For a homogeneous and isotropic Universe, they reduce to the Friedmann-Robertston-Walker (FRW) equations $$ \left( {\dot a \over a} \right)^2 - {k\over a^2} = {8 \pi G\over 3 } \rho \eqno\eq $$ $${\ddot a\over a} = - {4 \pi G\over 3} \, (\rho + 3 p) \, .\eqno\eq $$ These equations can be combined to yield the continuity equation (with Hubble constant $H = \dot a/a$) $$ \dot \rho = - 3 H (\rho + p) \, . \eqno\eq $$ The third key assumption of standard cosmology is that matter is described by an ideal gas with an equation of state $$ p = w \rho \, . \eqno\eq $$ For cold matter, pressure is negligible and hence $w = 0$. From (2.9) it follows that $$ \rho_m (t) \sim a^{-3} (t) \, , \eqno\eq $$ where $\rho_m$ is the energy density in cold matter. For radiation we have $w = {1/3}$ and hence it follows from (2.9) that $$ \rho_r (t) \sim a^{-4} (t) \, , \eqno\eq $$ $\rho_r (t)$ being the energy density in radiation. \section{Observational Pillars} The first observational pillar of standard cosmology is Hubble's redshift-distance relationship$^{18)}$ (Fig. 2) $$ z = H d \, , \eqno\eq $$ where $H$ is the present Hubble expansion constant, $d$ is the distance to a galaxy, and $z$ is its redshift $$ z \equiv {\lambda_0\over \lambda_e} - 1 \, , \eqno\eq $$ $\lambda_e (\lambda_0)$ being the wavelength of light at the time of emission (detection). \smallskip \epsfxsize=9cm \epsfbox{bfig2.eps} {\baselineskip=13pt \noindent{\bf Figure 2:} A recent redshift-distance plot of galaxies$^{19)}$. The distances are determined using the Tully-Fisher method. See Ref. 31 for a detailed discussion of the method and errors.} \medskip There is an easy intuitive derivation of this result. A wave in an expanding background will have a wavelength which increases as the scale factor $a(t)$. Hence for light emitted at time $t_e$ $$ z (t_e) = {a (t_0)\over{a (t_e)}} - 1 \, . \eqno\eq $$ For light emitted close to the present time we can Taylor expand the above result to obtain $$ z (t_e) \simeq {\dot a (t_0)\over{a (t_0)}} \, (t_0 - t_e) \simeq H (t_0) \, d \, . \eqno\eq $$ Equation (2.15) defines the cosmological redshift, which can be used as a measure of cosmic time. The second observational pillar of standard cosmology is the existence and black body nature of the CMB$^{20, 21)}$ To understand the connection$^{17)}$, consider matter in an expanding Universe. As we go backwards in time, the density of matter increases as $a^{-3} (t)$, and as a consequence the temperature grows. Above a temperature of 13.6 eV, atoms are ionized, and a bath of photons in thermal equilibrium must be present. Photons still scatter frequently below this temperature. At some time $t_{rec}$ the scattering length of a photon becomes longer than the Hubble radius. After that, photons travel without scattering. At $t_{rec}$, the distribution of photons is of black body type. A special feature of black body spectra is that the spectral shape is maintained even after $t_{rec}$. The only change is that the temperature redshifts $$ T(t) = {a (t_{rec})\over{a(t)}} \, T (t_{rec}) \, . \eqno\eq $$ Hence, the standard Big Bang model predicts a black body spectrum of photons with temperature $$ T_o = T_{rec} z (t_{rec})^{-1} \, , \eqno\eq $$ where $T_{rec}$ is the temperature at $t_{rec}$, determined by comparing the largest rate of scattering of photons below recombination, that due to Thomson scattering, with the Hubble expansion rate, yielding the result $$ T_{rec} \simeq 0.25 \, {\rm eV} \simeq 4000^\circ {\rm K} \, . \eqno\eq $$ The corresponding redshift is determined by measuring $T_o$. In 1965, Penzias and Wilson$^{22)}$ discovered this remnant black body radiation at a temperature of about 3$^\circ$ K. Since the spectrum peaks in the microwave region it is now called the CMB (cosmic microwave background). Recent satellite (COBE)$^{23)}$ and rocket$^{24)}$ experiments have confirmed the black body nature of the CMB to very high accuracy. The temperature is 2.73$^\circ$ K$= T_0$ which corresponds to $$ z (t_{rec}) = z_{rec} \sim 10^3 \, . \eqno\eq $$ Given the existence of the CMB, we know that matter has two components: dust (with energy density $\rho_m (t)$) and radiation (with density $\rho_r (t)$). At the present time $t_0$, $\rho_m (t) \gg \rho_r (t)$. The radiation energy density is determined by $T_0$, and the matter energy density can be estimated by analyzing the dynamics of galaxies and clusters and using the virial theorem. However, since by (2.11) and (2.12) $\rho_m (t) \sim a (t)^{-3}$ and $\rho_r (t) \sim a (t)^{-4}$, as we go back in time the fraction of energy density in radiation increases, and the two components become equal at a time $t_{eq}$, the time of equal matter and radiation. The corresponding redshift is $$ z_{eq} \simeq \Omega \, h^{-2}_{50} \, 10^4 \eqno\eq $$ where $$ \Omega = {\rho\over{\rho_c}} \, (t_0)\, , \eqno\eq $$ $\rho_c$ being the density for a spatially flat Universe (the critical density), and $h_{50}$ is the value of $H$ in units of 50 km s$^{-1}$ Mpc$^{-1}$. The time $t_{eq}$ is important for structure formation. As we will see in Section 4, it is only after $t_{eq}$ that perturbations on scales smaller than the Hubble radius $H^{-1} (t)$ can grow. Before then, the radiation pressure prevents growth. A temperature-time plot of the early Universe is sketched in Fig. 3, Note that $t_{eq} < t_{rec}$. \smallskip \epsfxsize=8cm \epsfbox{bfig3.eps} {\baselineskip=13pt \noindent{\bf Figure 3:} Temperature-time diagram of standard big bang cosmology. The present time, time of last scattering and time of equal matter and radiation are $t_0$, $t_{rec}$ and $t_{eq}$ respectively. The Universe is radiation-dominated before $t_{eq}$ (Region A) and matter-dominated in Region B. Before and after $t_{rec}$, respectively, the Universe was opaque and transparent, respectively, to microwave photons.} \medskip The third observational pillar of standard big bang cosmology concerns nucleosynthesis$^{25, 26)}$ - the production of light elements (heavy elements are formed in supernovae). Above a temperature of about $10^9 $ K, the nuclear interactions are sufficiently fast to prevent neutrons and protons from fusing. However, below that temperature, it is thermodynamically favorable for neutrons and protrons to fuse and form deuterium, helium 3, helium 4 and lithium 7 through a long and interconnected chain of reactions. The resulting light element abundances depend sensitively on the expansion rate of the Universe and on $\Omega_B$, the fraction of energy density $\rho_B$ at present in baryons relative to the critical density $\rho_c$. In Fig. 5, recent theoretical calculations$^{27)}$ of the abundances are shown and compared with observations. Demanding agreement with all abundances leaves only a narrow window $$ 3 \times 10^{-10} < \eta < 10^{-9} \, , \eqno\eq $$ where $\eta$ is the ratio of baryon number density $n_B$ to photon number density $n_\gamma$ $$ \eta = {n_b\over n_\gamma} \, . \eqno\eq $$ {}From (2.23), it follows that $\Omega_B$ is constrained: $$ 0.01 < \Omega_B h^2 < 0.035 \, . \eqno\eq $$ In particular, if the Universe is spatially flat and the cosmological constant is negligible, there must be nonbaryonic dark matter. We will return to the dark matter issue in Section 3. \smallskip \epsfxsize=10.5cm \epsfbox{bfig5.eps} {\baselineskip=13pt \noindent{\bf Figure 4:} Light element abundances as a function of the baryon to entropy ratio $\eta$ (from Ref. 27). The solid curves are the predictions of homogeneous big bang nucleosynthesis. The observational limits are indicated on the left vertical axis. Theory and observations are only consistent for a narrow range of values of $\eta$.} \medskip The final pillar of standard cosmology is the near isotropy of the CMB$^{16)}$. After subtracting the dipole anisotropy which is presumed to be due to the motion of the earth relative to the rest frame defined by the CMB, no anisotropies have been detected to a level of better than $10^{-4}$, i.e., the temperature difference $\delta T(\vartheta)$ between two beams pointing in directions in the sky separated by an angle $\vartheta$ (Fig. 4) satisfies $$ {\delta T (\vartheta)\over{\bar T}} < 10^{-4} \eqno\eq $$ on all angular scales $\vartheta$. Here $\bar T$ is the average temperature. \smallskip \epsfxsize=4.8cm \epsfbox{bfig4.eps} {\baselineskip=13pt \noindent{\bf Figure 5:} Sketch of a CMB anisotropy experiment. Two radio antennas with beam width $b$ collect microwave radiation from points in the sky separated by an angle $T$. The difference in beam intensities is measured.} \medskip Until recently, the isotropy of the CMB was the only observational support for the cosmological principle. Any inhomogeneities of the Universe on length scales comparable to the comoving Hubble radius at $t_{rec}$ and larger would generate temperature anisotropies by a mechanism discussed in detail in Section 7. Hence, the near isotropy of the CMB implies that density fluctuations on large scales must have been very small in the early Universe. To summarize, the observational pillars of standard cosmology are Hubble's redshift-distance relation, the existence and black body nature of the CMB, primordial nucleosynthesis, and the isotropy of the CMB. Note, in particular, that no tests of big bang cosmology say anything about the evolution of the Universe before the time of nucleosynthesis. Note, also, that not all astronomers accept the above observations as support of the Big Bang model. For a recent criticism see Ref. 28 (and Ref. 29 for a reply to the criticism). \section{Problems} Standard Big Bang cosmology is faced with several important problems. Only one of these, the age problem, is a potential conflict with observations. The others which I will focus on here -- the homogeneity, flatness and formation of structure problems (see e.g. Ref. 30) -- are questions which have no answer within the theory and are therefore the main motivation for the new cosmological models which will be discussed in the rest of these lectures. {}From the FRW equations (2.7) and (2.8) it is easy to calculate the age of the Universe, given the expansion law (2.11) which holds throughout most of the history of the Universe. For a spatially flat Universe, the age $\tau$ depends on the expansion rate $H$, i.e. on the constant $h$ which is in the range $0.4 < h < 1$: $$ \tau \simeq {7\over h} 10^9 {\rm yr} \, . \eqno\eq $$ Globular cluster ages have been estimated to lie in the range $12 - 18 \times 10^9$ yr. Thus, theory and observations are only consistent if $h < 0.55$ (see e.g. Ref. 31). In an open Universe the problem is less severe. Recent observations have not led to a decrease in the uncertainty in the value of $h$. Observations by the Hubble space telescope$^{209)}$ and on supernovae observations$^{210)}$ indicate a fairly large value ($h \simeq 0.8$), but direct measurements based on the Sunyaev-Zeldovich effect and using more distant galaxy clusters yield$^{211)}$ a small value ($h \simeq 0.5$). Modern cosmological models do not add any insight into the age problem since they only modify the evolution of the Universe at very early times $t \ll t_{eq}$. The final three problems mentioned above, the homogeneity, flatness and formation of structure problems, provided a lot of the motivation for the development of the inflationary Universe scenario$^{30)}$ and will hence be discussed in detail. The horizon problem is illustrated in Fig. 6. As is sketched, the comoving region $\ell_p (t_{rec})$ over which the CMB is observed to be homogeneous to better than one part in $10^4$ is much larger than the comoving forward light cone $\ell_f (t_{rec})$ at $t_{rec}$, which is the maximal distance over which microphysical forces could have caused the homogeneity: $$ \ell_p (t_{rec}) = \int\limits^{t_0}_{t_{rec}} dt \, a^{-1} (t) \simeq 3 \, t_0 \left(1 - \left({t_{rec}\over t_0} \right)^{1/3} \right) \eqno\eq $$ $$ \ell_f (t_{rec}) \int\limits^{t_{rec}}_0 dt \, a^{-1} (t) \simeq 3 \, t^{2/3}_0 \, t^{1/3}_{rec} \, . \eqno\eq $$ {}From the above equations it is obvious that $\ell_p (t_{rec}) \gg \ell_f (t_{rec})$. Hence, standard cosmology cannot explain the observed isotropy of the CMB. \smallskip \epsfxsize=6.5cm \epsfbox{bfig6.eps} {\baselineskip=13pt \noindent {\bf Figure 6:} A space-time diagram (physical distance $x_p$ versus time $t$) illustrating the homogeneity problem: the past light cone $\ell_p (t)$ at the time $t_{rec}$ of last scattering is much larger than the forward light cone $\ell_f (t)$ at $t_{rec}$.} \medskip In standard cosmology and in an expanding Universe, $\Omega = 1$ is an unstable fixed point. This can be seen as follows. For a spatially flat Universe $(\Omega = 1)$ $$ H^2 = {8 \pi G\over 3} \, \rho_c \, , \eqno\eq $$ whereas for a nonflat Universe $$ H^2 + \varepsilon \, T^2 = {8 \pi G\over 3} \, \rho \, , \eqno\eq $$ with $$ \varepsilon = {k\over{(aT)^2}} \, . \eqno\eq $$ The quantity $\varepsilon$ is proportional to $s^{-2/3}$, where $s$ is the entropy density. Hence, in standard cosmology, $\varepsilon$ is constant. Combining (2.30) and (2.31) gives $$ {\rho - \rho_c\over \rho_c} = {3\over{8 \pi G}} \, {\varepsilon T^2\over \rho_c} \sim T^{-2} \, . \eqno\eq $$ Thus, as the temperature decreases, $\Omega - 1$ increases. In fact, in order to explain the present small value of $\Omega - 1 \sim {\cal O} (1)$, the initial energy density had to be extremely close to critical density. For example, at $T = 10^{15}$ GeV, (2.33) implies $$ {\rho - \rho_c\over \rho_c} \sim 10^{-50} \, . \eqno\eq $$ What is the origin of these fine tuned initial conditions? This is the flatness problem of standard cosmology. The last problem of the standard cosmological model I will mention is the ``formation of structure problem." Observations indicate that galaxies and even clusters of galaxies have nonrandom correlations on scales larger than 50 Mpc (see e.g. Ref. 14). This scale is comparable to the comoving horizon at $t_{eq}$. Thus, if the initial density perturbations were produced much before $t_{eq}$, the correlations cannot be explained by a causal mechanism. Gravity alone is, in general, too weak to build up correlations on the scale of clusters after $t_{eq}$ (see, however, the explosion scenario of Ref. 32). Hence, the two questions of what generates the primordial density perturbations and what causes the observed correlations, do not have an answer in the context of standard cosmology. This problem is illustrated by Fig. 7. \smallskip \epsfxsize=6.5cm \epsfbox{bfig7.eps} {\baselineskip=13pt \noindent {\bf Figure 7:} A sketch (conformal separation vs. time) of the formation of structure problem: the comoving separation $d_c$ between two clusters is larger than the forward light cone at time $t_{eq}$.} \medskip Finally, let us address the cosmological constant problem. All known symmetries of nature and principles of general relativity allow for the presence of a term in the Einstein equations which acts like matter with energy $\Lambda$ and pressure $-\Lambda$, i.e., with an equation of state $p = - \rho$. If it is not to dominate the present expansion rate of the Universe, the cosmological constant $\Lambda$ must be very small $$ \Lambda < 3 H^2_0 \sim 10^{-83} {\rm GeV}^2 \, . \eqno\eq $$ On dimensional grounds, we would expect $\Lambda$ to be of the order $m_{pl}^2 \sim 10^{38} \, {\rm GeV}^2$. Thus, the cosmological constant is about 140 orders of magnitude smaller than what we would expect it to be (for recent reviews of the cosmological constant problem, see Ref. 33). As we will see, modern cosmology does not address the cosmological constant problem. If anything, the problem will manifest itself in a more apparent manner. For some recent ideas on how infrared effects in field theory might solve the cosmological constant problem see Ref. 212. Due to the formation of structure problem, there can be no causal physical theory for the origin of structure (with nontrivial spatial correlations) in the Universe in the context of the Standard Big Bang theory. The main breakthrough of modern cosmology is that it provides solutions to this problem. The key to understanding this breakthrough in cosmology is the realization of the internal inconsistency of the standard picture when extrapolated to times much before nucleosynthesis. Standard cosmology is based on the assumption that matter continues to be described by an ideal radiation gas to arbitrarily high temperatures. This is clearly in contrast to what nuclear and particle physics tells us. As we go backwards in time towards the Big Bang, nuclear physics and eventually particle physics effects will take over. To describe matter correctly, a quantum field theoretic description must be used. Note, however, that at a fundamental level there is an inconsistency if matter is described quantum mechanically while maintaining a classical description of gravity. Hence, we cannot hope that any of the present cosmological theories will be the ultimate theory. \chapter{New Cosmology and Structure Formation} The goal of this section is to present an overview of what can be gained if we go beyond standard cosmology and allow matter to be described in terms of concepts from particle physics. Detailed discussions of the models will be given in later sections. \section{The Inflationary Unvierse} The idea of inflation$^{30)}$ is very simple. We assume there is a time interval beginning at $t_i$ and ending at $t_R$ (the ``reheating time") during which the Universe is exponentially expanding, i.e., $$ a (t) \sim e^{Ht}, \>\>\>\>\> t \epsilon \, [ t_i , \, t_R] \eqno\eq $$ with constant Hubble expansion parameter $H$. Such a period is called ``de Sitter" or ``inflationary." The success of Big Bang nucleosynthesis sets an upper limit to the time of reheating: $$ t_R \ll t_{NS} \, , \eqno\eq $$ $t_{NS}$ being the time of nucleosynthesis. \medskip \epsfxsize=7cm \epsfbox{bfig8.eps} {\baselineskip=13pt \noindent{\bf Figure 8:} The phases of an inflationary Universe. The times $t_i$ and $t_R$ denote the beginning and end of inflation, respectively. In some models of inflation, there is no initial radiation domintated FRW period. Rather, the classical space-time emerges directly in an inflationary state from some initial quantum gravity state.} \medskip The phases of an inflationary Universe are sketched in Fig. 8. Before the onset of inflation there are no constraints on the state of the Universe. In some models a classical space-time emerges immediately in an inflationary state, in others there is an initial radiation dominated FRW period. Our sketch applies to the second case. After $t_R$, the Universe is very hot and dense, and the subsequent evolution is as in standard cosmology. During the inflationary phase, the number density of any particles initially in thermal equilibrium at $t = t_i$ decays exponentially. Hence, the matter temperature $T_m (t)$ also decays exponentially. At $t = t_R$, all of the energy which is responsible for inflation (see later) is released as thermal energy. This is a nonadiabatic process during which the entropy increases by a large factor. The temperature-time evolution in an inflationary Universe is depicted in Fig. 9. \smallskip \epsfxsize=6cm \epsfbox{bfig9.eps} {\baselineskip=13pt \noindent{\bf Figure 9:} The time dependence of matter temperature in an inflationary Universe. During the period of exponential expansion, the temperature decreases exponentially. At the end of inflation the energy density of the scalar field responsible for inflation is transferred to ordinary matter. This leads to reheating. The critical temperature $T_c$ is the temperature at which the initial matter thermal energy density becomes less than the scalar field energy density (see Chapter 5).} \medskip Fig. 10 is a sketch of how a period of inflation can solve the homogeneity problem. $\Delta t = t_R - t_i$ is the period of inflation. During inflation, the forward light cone increases exponentially compared to a model without inflation, whereas the past light cone is not affected for $t \geq t_R$. Hence, provided $\Delta t$ is sufficiently large, $\ell_f (t_R)$ will be greater than $\ell_p (t_R)$. The condition on $\Delta t$ depends on the temperature $T_R$ corresponding to time $t_R$, the temperature of reheating. Demanding that $\ell_f (t_R) > \ell_p (t_R)$ we find, using the analogs of (2.28) and (2.29), the following criterion $$ e^{\Delta t H} \geq \, {\ell_p (t_R)\over{\ell_f (t_R) }} \simeq \left( {t_0\over t_R} \right)^{1/2} = \, \left({T_R\over T_0} \right) \sim 10^{27} \eqno\eq $$ for $T_R \sim 10^{14}$GeV and $T_0 \sim 10^{-13}$GeV (the present microwave background temperature). Thus, in order to solve the homogeneity problem, a period of inflation with $$ \Delta t \gg 50 \, H^{-1} \eqno\eq $$ is required. {\baselineskip=13pt \noindent{\bf Figure 10:} Sketch (physical coordinates vs. time) of the solution of the homogeneity problem. During inflation, the forward light cone $l_f(t)$ is expanded exponentially when measured in physical coordinates. Hence, it does not require many e-foldings of inflation in order that $l_f(t)$ becomes larger than the past light cone at the time of last scattering. The dashed line is the forward light cone without inflation.} \medskip Inflation also can solve the flatness problem$^{34, 30)}$ The key point is that the entropy density $s$ is no longer constant. As will be explained later, the temperatures at $t_i$ and $t_R$ are essentially equal. Hence, the entropy increases during inflation by a factor $\exp (3 H \Delta t)$. Thus, $\epsilon$ decreases by a factor of $\exp (-2 H \Delta t)$. With the numbers used in (3.3): $$ \epsilon_{\rm after} \sim 10^{-54} \, \epsilon_{\rm before} \, . \eqno\eq $$ Hence, $(\rho - \rho_c) / \rho$ can be of order 1 both at $t_i$ and at the present time. In fact, if inflation occurs at all, the theory then predicts that at the present time $\Omega = 1$ to a high accuracy (now $\Omega < 1$ would require special initial conditions). What was said above can be rephrased geometrically: during inflation, the curvature radius of the Universe -- measured on a fixed physical scale -- increases exponentially. Thus, a piece of space looks essentially flat after inflation even if it had measurable curvature before. Most importantly, inflation provides a mechanism which in a casual way generates the primordial perturbations required for galaxies, clusters and even larger objects. In inflationary Universe models, the Hubble radius (``apparent" horizon), $3t$, and the ``actual" horizon (the forward light cone) do not coincide at late times. Provided (3.3) is satisfied, then (as sketched in Fig. 11) all scales within our apparent horizon were inside the actual horizon since $t_i$. Thus, it is in principle possible to have a casual generation mechanism for perturbations$^{35-38)}$. {\baselineskip=13pt \noindent{\bf Figure 11:} A sketch (physical coordinates vs. time of the solution of the formation of structure problem. Provided that the period of inflation is sufficiently long, the separation $d_c$ between two galaxy clusters is at all times smaller than the forward light cone. The dashed line indicates the Hubble radius. Note that $d_c$ starts out smaller than the Hubble radius, crosses it during the de Sitter period, and then reenters it at late times.} \medskip The generation of perturbations is supposed to be due to a causal microphysical process. Such processes can only act coherently on length scales smaller than the Hubble radius $\ell_H (t)$ where $$ \ell_H (t) = H^{-1} (t) \, . \eqno\eq $$ A heuristic way to understand the meaning of $\ell_H (t)$ is to realize that it is the distance which light (and hence the maximal distance any causal effects) can propagate in one expansion time: $$ \ell_H (t) \sim a (t) \int\limits_{t}^{t+H^{-1} (t)} a (t^\prime)^{-1} \, dt^\prime \, . \eqno\eq $$ In Section 5 a more mathematical justification for the definition and role of $\ell_H (t)$ will be given. As will be discussed in Section 5, the density perturbations produced during inflation are due to quantum fluctuations in the matter and gravitational fields$^{36, 37)}$. The amplitude of these inhomogeneities corresponds to a tempertuare $T_H$ $$ T_H \sim H \, , \eqno\eq $$ the Hawking temperature of the de Sitter phase. This implies that at all times $t$ during inflation, perturbations with a fixed physical wavelength $\sim H^{-1}$ will be produced. Subsequently, the length of the waves is streched with the expansion of space, and soon becomes larger than the Hubble radius. The phases of the inhomogeneities are random. Thus, the inflationary Universe scenario predicts perturbations on all scales ranging from the comoving Hubble radius at the beginning of inflation to the corresponding quantity at the time of reheating. In particular, provided that inflation lasts sufficiently long (see (3.4)), perturbations on scales of galaxies and beyond will be generated. Note, however, that it is very dangerous to interpret de Sitter Hawking radiation as thermal radiation. In fact, the euation of state of this ``radiation" is not thermal$^{213)}$. Now that the reader is (hopefully) convinced that inflation is a beautiful idea, the question arises how to realize this scenario. The initial hope was that the same scalar fields (Higgs fields) which particle physicists introduce in order to spontaneously break the internal symmetries of their field theory models would lead to inflation. This hope was based on the fact that the energy density $\rho$ and pressure $p$ of a real scalar field $\varphi (\underline{x}, t)$ are given by $$ \eqalign{\rho (\varphi) & = {1\over 2} \, \dot \varphi^2 + {1\over 2} \, (\nabla \varphi)^2 + V (\varphi) \cr p (\varphi) & = {1\over 2} \dot \varphi^2 - {1\over 6} (\nabla \varphi)^2 - V (\varphi) \, .} \eqno\eq $$ Thus, provided that at some initial time $t_i$ $$ \dot \varphi (\underline{x}, \, t_i) = \nabla \varphi (\underline{x}_i \, t_i) = 0 \eqno\eq $$ and $$ V (\varphi (\underline{x}_i \, t_i) ) > 0 \, , \eqno\eq $$ the equation of state of matter will read $$ p = - \rho \eqno\eq $$ and, from the FRW equations it will follow that $$ a (t) = e^{tH} \> , \> H^2 = \, {8 \pi G\over 3} \, V (\varphi) \, . \eqno\eq $$ The next question is how to realize the required initial conditions (3.10) and to maintain the key constraints $$ \dot \varphi^2 \ll V (\varphi) \> , \> (\nabla \varphi)^2 \ll V (\varphi) \eqno\eq $$ for sufficiently long (see (3.4)). This typically requires potentials which are very flat near $\varphi (\underline{x}, \, t_i)$. Worse yet, the parameters of the potential $V (\varphi)$ must be chosen such that the final amplitude of density perturbations is sufficiently small to agree with the limits on CMB anisotropies. As we will demonstrate in Section 5, these conditions impose severe constraints on the constants which appear in $V (\varphi)$. In light of these difficulties it is important to keep in mind that inflation can also be generated by modifying gravity at high curvatures (see e.g., Refs. 39-41). It is also wise to investigate alternative theories of structure formation which do not rely on inflation. \section{Topological Defect Models} According to particle physics theories, matter at high energies and temperatures must be described in terms of fields. Gauge symmetries have proved to be extremely useful in describing the standard model of particle physics, according to which at high energies the laws of nature are invariant under a nonabelian group $G$ of internal symmetry transformations $$ G = {\rm SU} (3)_c \times {\rm SU} (2)_L \times U(1)_Y \eqno\eq $$ which at a temperature of about 200 MeV is spontaneously broken down to $$ G^\prime = {\rm SU} (3)_c \times {\rm U} (1) \, . \eqno\eq $$ The subscript on the SU(3) subgroup indicates that it is the color symmetry group of the strong interactions, ${\rm SU} (2)_L \times $ U(1)$_Y$ is the Glashow-Weinberg-Salam (WS) model of weak and electromagnetic interactions, the subscripts $L$ and $Y$ denoting left handedness and hypercharge respectively. At low energies, the WS model spontaneously breaks to the U(1) subgroup of electromagnetism. Spontaneous symmetry breaking is induced by an order parameter $\varphi$ taking on a nontrivial expectation value $< \varphi >$ below a certain temperture $T_c$. In some particle physics models, $\varphi$ is a fundamental scalar field in a nontrivial representation of the gauge group $G$ which is broken. However, $\varphi$ could also be a fermion condensate, as in the BCS theory of superconductivity. The transition taking place at $T = T_c$ is a phase transition and $T_c$ is called the critical temperature. From condensed matter physics it is well known that in many cases topological defects form during phase transitions, particularly if the transition rate is fast on a scale compared to the system size. When cooling a metal, defects in the crystal configuration will be frozen in; during a temperature quench of $^4$He, thin vortex tubes of the normal phase are trapped in the superfluid; and analogously in a temperature quench of a superconductor, flux lines are trapped in a surrounding sea of the superconducting Meissner phase (see Fig. 12 for an example$^{42)}$ of defect formation). {\baselineskip=13pt \noindent{\bf Figure 12:} A simulation of defect formation in the $2 + 1$ dimensional Abelian Higgs model$^{42)}$ in an expanding background. The total energy density is plotted against the two spatial coordinates. The initial conditions were specified by thermal initial conditions with random phases of the order parameter on Hubble scales. At later times, the thermal noise has redshifted away, leaving behind trapped energy density in vortices.} \medskip In cosmology, the rate at which the phase transition proceeds is given by the expansion rate of the Universe. Hence, topological defects will inevitably be produced in a cosmological phase transition$^{12)}$, provided the underlying particle physics model allows such defects. Topological defects can be point-like (monopoles), string-like (cosmic strings)$^{43)}$ or planar (domain walls), depending on the particle physics model (see Section 6). Also of importance are textures$^{44, 45)}$, point defects in space-time. Topological defects represent regions in space with trapped energy density. These regions of surplus energy can act as seeds for structure formation as is illustrated in Fig. 13. For point-like defects, the force which causes clustering about the seed can be understood using Newtonian gravity. The process is called gravitational accretion. For precise calculations (and in the case of other defects), general relativistic effects must be taken into account (see Section 6). {\baselineskip=13pt \noindent{\bf Figure 13:} Sketch of the basic gravitational accretion mechanism. The topological defect (in this case a cosmic string loop) is a configuration of trapped energy density. This excess density produces a Newtonian gravitational attractive force on the surrounding matter.} \medskip No stable topological defects arise in the breaking of the WS model. However, there is good evidence for phase transitions at very high energies. The coupling constants of SU(3)$_c$, SU(2) and U(1) are seen to converge at an energy scale $\eta$ of about $$ \eta \sim 10^{16} \, {\rm GeV} \, . \eqno\eq $$ It is therefore not unreasonable to speculate that the standard model results from the breaking of a larger symmetry group $G_0$ at a scale $\eta$. A large class of unified gauge theories$^{46)}$ based on a symmetry breaking $$ G_0 \longrightarrow G = {\rm SU} (3) \times {\rm SU}(2) \times {\rm U}(1) \eqno\eq $$ admit topological defects, many theories have cosmic string solutions. Topological defect models of structure formation will be discussed in detail in Section 6. Here I will briefly point out by which mechanism correlations on all cosmological scales are induced. To be concrete, I consider the cosmic string model$^{47, 48)}$. Cosmic strings are one-dimensional defects without ends. Hence, they must be either infinite in length or else closed loops. The fact that at the time of the phase transition $t_c$, the order parameter has random phases on scales larger than the initial correlation length implies that at $t_c$ a random walk-like network of infinite strings will form$^{12)}$. This implies nontrivial correlations of structures seeded by these strings on all scales larger than the initial correlation length. \section{Need for Dark Matter} At this point we have illustrated two classes of mechanisms by which structure formation in the Universe can be seeded: quantum fluctuations during a period of inflation, and topological defects. To completely specify a theory of structure formation, however, we must also specify what the ``dark matter" which dominates the energy density of the Universe today consists of (see e.g. Ref. 49). The evidence for dark matter has been accumulating over the past decade. Measurements of galaxy velocity rotation curves indicate that a large fraction of the mass of a galaxy does not shine. In Fig. 14, the measured rotation velocity $v$ is plotted as a function of the distance $r$ from the center of the galaxy. This velocity is compared to the velocity $\hat v (r)$ which results from the virial theorem, assuming that light traces mass (the luminosity profile is given in the upper frame). The data is for the spiral galaxies NGC2403 and NGC3198$^{50)}$. The comparison shows that the mass of these galaxies extends significantly beyond the visible radius and that -- assuming Newtonian gravity is applicable -- a large fraction of the mass of a spiral galaxy must be dark. {\baselineskip=13pt \noindent{\bf Figure 14:} Velocity rotation curves for two galaxies (from Ref. 50). The upper panel shows the luminosity of the galaxy as a function of the distance from the center, the lower panel presents the velocity (vertical axis, in units of $km s^{-1}$) as a function of the same distance. The solid curve is the velocity inferred from the observed luminosity curve, using the virial theorem, the dotted curves are the observational results. The fact that the velocity rotation curves remain constant beyond the visible radius of the galaxy is strong evidence for galactic dark matter.} \medskip The fraction of matter that shines can be expressed as a fraction $\Omega_{\rm lum}$ of the critical density $\rho_c$. The present estimates give $$ \Omega_{\rm lum} \ll 0.01 \, , \eqno\eq $$ whereas the fraction $\Omega_g$ of mass in galaxies is much larger$^{51)}$ $$ \Omega_g \sim 0.03 \, . \eqno\eq $$ It is also possible to estimate the fraction $\Omega_{cl}$ of mass which is gravitationally bound in clusters. Current estimates using the virial theorem give$^{51)}$ $$ \Omega_{cl} \sim 0.1 - 0.2 \eqno\eq $$ (this comes mainly from studying the infall of galaxies towards the Virgo cluster). Finally, the amount $\Omega_{LSS}$ of mass in large-scale structures can be estimated by measuring the large-scale peculiar velocities and inferring the mass required to generate such velocities. Current estimates give$^{52, 53)}$ $$ \Omega_{LSS} = 0.8 \pm 0.5 \, . \eqno\eq $$ As mentioned in Section 2, nucleosynthesis provides independent limits on the fraction $\Omega_B$ of mass in baryons (see (2.25)). Comparing (3.19-3.22) with (2.25) we conclude: \item{i)} There must be baryonic dark matter. In fact, most of the dark matter in galaxies and/or clusters could be baryonic, and some of this baryonic dark matter may have recently been discovered by gravitational microlensing. \item{ii)} If $\Omega = 1$, then most of the matter in the Universe consists of nonbaryonic dark matter. In fact, there is increasing evidence (see (3.22)) that nonbaryonic dark matter must exist independent of the theoretical prejudice for $\Omega = 1$. The dark matter in the Universe is visible only through its gravitational effects. Hence, nonbaryonic dark matter candidates can be divided into two classes, cold dark matter (CDM) and hot dark matter (HDM). CDM particles are cold, i.e., their peculiar velocity $v$ is negligible at the time $t_{eq}$ when structure formation begins: $$ v (t_{eq}) \ll 1 \, . \eqno\eq $$ Candidates for CDM include the axion (coherent oscillations of a low mass scalar field) and neutralinos (the lightest stable supersymmetric particle, which must be neutral). HDM particles are relativistic at $t_{eq}$: $$ v (t_{eq}) \sim 1 \, . \eqno\eq $$ The prime candidate is a $25 h^{+2}_{50}$eV tau neutrino. Note that this mass is well within the experimental bounds for the tau neutrino mass, and also that many particle physics models -- in particular those which lead to neutrino oscillations -- predict masses of this order of magnitude. \section{Survey of Models} Any theory of structure formation must specify both the source of fluctuations and the composition of the dark matter. The reader is warned that the model called the ``CDM Model" is a model with CDM {\bf AND} perturbations generated by quantum fluctuations during a hypothetical period of inflation. Inflation-based models were the first to be considered in quantitative detail, initially assuming a HDM-dominated Universe. Almost immediately, however, contradictions with basic observations appeared$^{54)}$ (see, however, Ref. 214 for an opposing point of view). The problem of HDM-based inflationary models is related to neutrino free streaming$^{55)}$. The primordial perturbations in this theory are dark matter fluctuations, but because of the large velocity of the dark matter particles, the inhomogeneities are washed out on all scales below the neutrino free streaming length $\lambda^c_j (t)$, $$ \lambda^c_j (t) \sim v (t) z (t) t \, , \eqno\eq $$ which is the comoving distance the particles move in one Hubble expansion time. Since the neutrino velocity $v(t)$ and the redshift $z(t)$ both scale as $a(t)^{-1}$, the free streaming length decreases as $$ \lambda^c_j (t) \sim t^{-1/3} \eqno\eq $$ after $t_{eq}$ (before $t_{eq}$ the radiation pressure dominates). Hence, in an inflationary HDM model all perturbations on scales $\lambda$ smaller than the maximal value of $\lambda^c_j (t)$ are erased. The critical scale $\lambda^{\rm max}_j$ is given by the value of $\lambda^c_j (t)$ at the time when the neutrinos become non-relativistic which is in turn determined by the neutrino mass $m_{\nu}$. The result is $$ \lambda^{\rm max}_j \simeq 30 \, {\rm Mpc} \, \left({{m_{\nu}} \over {25 {\rm eV}}} \right)^{-2} \, , \eqno\eq $$ a scale much larger than the mean separation of galaxies and clusters. Since we observe galaxies outside of large-scale structures, this model is in blatant disagreement with observations. Inflation-based models are hence only viable if (at least a substantial fraction of) the dark matter is cold. Such models have become known as ``CDM models", and are to a first approximation rather successful at predicting the clustering properties of galaxies and galaxy clusters$^{56)}$. There are many parameters in CDM models: the amplitude of the density perturbations, the power of the spectrum (see Section 4), the value of $\Omega$, the fraction $\Omega_B$ of baryons, to mention some of the main ones. It is also possible to add a small fraction $\Omega_v$ of hot dark matter (yielding a class of so-called ``Mixed Dark Matter" models). Topological defect models were first developed in the context of CDM. Theories based on cosmic strings$^{57-59)}$ or on global textures$^{45, 60)}$ have also been fairly successful in explaining observations (again to a first approximation). It is important to note that if perturbations are seeded by long-lived topological defects (e.g., cosmic strings), then the above arguments against hot dark matter disappear$^{61, 62)}$. The seed perturbations can survive neutrino free streaming as long as the seeds remain present for many Hubble expansion times. If we consider a comoving scale $\lambda$ much smaller than $\lambda^{\rm max}_j$ of (3.27), then a dark matter perturbation will begin to grow about the seed fluctuations at a time $t(\lambda)$ when $$ \lambda^c_j (t (\lambda)) = \lambda \, . \eqno\eq $$ Cosmic string based hot dark matter models have also been successful at explaining the qualitative features of observations$^{63)}$. \chapter{Basics of Structure Formation} In the structure formation models mentioned in the previous section, small amplitude seed perturbations are predicted to arise due to particle physics effects in the very early Universe. They then grow by gravitational instability to produce the cosmological structures we observe today. In order to be able to make the connection between particle physics and observations, it is important to understand the gravitational evolution of fluctuations. This section will introduce the basic concepts of this topic. We begin, however, with an overview of some of the relevant data. \section{Survey of Data} It is length scales corresponding to galaxies and larger which are of greatest interest in cosmology when attempting to find an imprint of the primordial fluctuations produced by particle physics. On these scales, gravitational effects are assumed to be dominant, and the fluctuations are not too far from the linear regime. On smaller scales, nonlinear gravitational and hydrodynamical effects determine the final state and mask the initial perturbations. To set the scales, consider the mean separation of galaxies, which is about 5$h^{-1}$ Mpc$^{64)}$, and that of Abell clusters which is around 25$h^{-1}$ Mpc$^{65)}$. The largest coherent structures seen in current redshift surveys have a length of about 100$h^{-1}$ Mpc, the recent detections of CMB anisotropies probe the density field on length scales of about $10^3 h^{-1}$ Mpc, and the present horizon corresponds to a distance of about $3 \cdot 10^3 h^{-1}$ Mpc. Galaxies are gravitationally bound systems containing billions of stars. They are non-randomly distributed in space. A quantitative measure of this non-randomness is the ``two-point correlation function" $\xi_2 (r)$ which gives the excess probability of finding a galaxy at a distance $r$ from a given galaxy: $$ \xi_2 (r) = < \, {n (r) - n_0\over n_0} \, > \, . \eqno\eq $$ Here, $n_0$ is the average number density of galaxies, and $n(r)$ is the density of galaxies a distance $r$ from a given one. The pointed braces stand for ensemble averaging. \smallskip \epsfxsize=13cm \epsfbox{bfig15.eps} {\baselineskip=13pt \noindent{\bf Figure 15:} Recent observational results for the two point correlation function of IRAS galaxies, in both a volume limited subsample and a complete galaxy sample (see Ref. 66 which explains the meaning of the measure $J_3(r)$).} \medskip Recent observational results from a redshift survey of IRAS (infrared) galaxies yields reasonable agreement$^{66)}$ with a form (see Fig. 15) $$ \xi_2 (r) \simeq \left({r_0\over r} \right)^\gamma \eqno\eq $$ with scaling length $r_0 \simeq 5 h^{-1}$ Mpc and power $\gamma \simeq 1.8$. A theory of structure formation must explain both the amplitude and the slope of this correlation function. Galaxies do not all have the same mass. There are more smaller galaxies than large ones (our galaxy is a large one). The distribution of galaxy masses is given by the ``galaxy mass function" $n(M)$, where $n (M) dM$ is the number density of galaxies in the mass range $[M, \, M+dM]$. Since we can only measure luminosity but not mass, the observable measure of the galaxy distribution is $\phi (L)$, the ``galaxy luminosity function." Now, $\phi (L) dL$ is the number density of galaxies with luminosity $L$ in the interval $[L, \, L+dL]$. Recent results for $\phi (L)$ from the IRAS galaxy survey are reproduced in Fig. 16$^{67)}$. \smallskip \epsfxsize=11cm \epsfbox{bfig16.eps} {\baselineskip=13pt \noindent{\bf Figure 16:} The luminosity function of IRAS galaxies. The vertical axis is $\phi (L)$, the horizontal axis is luminosity $L$ in units of solar luminosity. The solid curve represents the data, the dashed curve is a fit to a theoretical model (see Ref. 67).} \medskip Theories must also be able to explain the internal mass distribution of galaxies which can be inferred from the galaxy rotation curves (see Fig. 14). According to the virial theorem, the velocity $v(r)$ at a distance $r$ from the center of a galaxy is determined by the mass $M(r)$ inside of $r$: $$ {mv^2\over r} = G \, {m M(r)\over r^2} \eqno\eq $$ where $m$ is a test mass. Hence $$ v (r) = \left(G \, {M (r)\over r} \right)^{1/2} \, . \eqno\eq $$ A constant rotation velocity implies $M(r) \sim r$ and hence a density profile $\rho (r)$ of $$ \rho (r) \sim r^{-2} \, . \eqno\eq $$ This condition puts constraints on the possible composition of the galactic dark matter. An Abell cluster$^{68)}$ is a region in space with greater than fifty bright galaxies in a sphere of radius $1.5 h^{-1}$ Mpc, i.e., a region with a very high overdensity of galaxies. Observations indicate that Abell clusters are not distributed randomly in space. The cluster two point correlation function $\xi_c (r)$ has a form similar to (4.2)$^{65, 69)}$: $$ \xi_c (r) \simeq \left({r_0\over r} \right)^\gamma \eqno\eq $$ with $r_0 \simeq 15 h^{-1}$ Mpc and $\gamma \simeq 2$ (see Fig. 17 which is taken from a recent analysis of rich clusters of galaxies selected from the APM Galaxy Survey$^{69)}$). \smallskip \epsfxsize=11cm \epsfbox{bfig17.eps} {\baselineskip=13pt \noindent{\bf Figure 17:} The two point correlation function $\xi_c$ (denoted by $\xi_{cc}$ in the figure) of clusters of galaxies drawn from the APM galaxy survey as a function of their separation $s$. Results for two different samples of clusters are given.} \medskip There are several remarkable features about the clustering of galaxies and galaxy clusters. First, there is evidence for universality of the functional form of $\xi (r)$; its slope is about 2 for both populations. Secondly, relative to their respective mean separations, galaxies are more clustered than galaxy clusters. This can be explained by the action of gravity. Gravity has had longer to act on the scales of galaxies than on that of clusters, and has hence amplified the galaxy correlation function relative to that of clusters. There is a wide spread of cluster masses which can be described by the cluster multiplicity function $\Phi_c (n)$, where $\Phi_c (n) dn$ is the number density of clusters containing between $n$ and $n+dn$ galaxies. Fig. 18 is a sketch of the observed cluster multiplicity function taken from Ref. 70. Note that the cluster mass function inferred from Fig. 18 and the galaxy mass function deduced from the galaxy luminosity function of Fig. 16 match up quite well at a mass of about $10^{12} M_\odot$. Below this mass, the objects are well defined dynamical entities whereas for larger masses they are composed of fragments. A reason for the difference may be due to the fact that clouds of more than $10^{12} M_\odot$ cannot cool without fragmenting$^{71)}$. \smallskip \epsfxsize=10.5cm \epsfbox{bfig18.eps} {\baselineskip=13pt \noindent{\bf Figure 18:} The cluster multiplicity function $\Phi_c (n)$ (denoted as $n$ on the figure) as a function of the number $n$ (horizontal axis).} \medskip On scales larger than galaxy clusters there is not at present a clear mathematical description of structure. Many galaxy redshift surveys have discovered coherent filamentary and planar structures and voids on scales of up to $100 h^{-1}$ Mpc$^{14, 72-75)}$. For example, the astronomers working on the ``Center for Astrophysics" redshift survey$^{14)}$ have analyzed many adjacent slices of the northern celestial sphere. For all galaxies above a limiting magnitude of 15.5 they measured the redshifts $z$. Fig. 19 is a sketch of redshift versus angle $\alpha$ in the sky for one slice. The second direction in the sky has been projected onto the $\alpha -z$ plane. The most prominent feature is the band of galaxies at a distance of about $100h^{-1}$ Mpc. This band also appears in neighboring slices and is therefore presumably part of a planar density enhancement of comoving planar size of at least $(50 \times 100) \times h^{-2}$ Mpc$^2$. This structure is often called the ``great wall." It is a challenge for theories of structure formation to explain both the observed scale and topology of the galaxy distribution. \smallskip \epsfxsize=9cm \epsfbox{bfig19.eps} {\baselineskip=13pt \noindent{\bf Figure 19:} Results from the CFA redshift survey. Radial distance gives the redshift of galaxies, the angular distance corresponds to right ascension. The results from several slices of the sky (at different declinations) have been projected into the same cone.} \medskip Until 1992 there was little evidence for any convergence of the galaxy distribution towards homogeneity. Each new survey led to the discovery of new coherent structures in the Universe on a scale comparable to that of the survey. In 1992, preliminary results of a much deeper redshift survey were announced$^{15)}$ which for the first time found no new coherent structures on scales larger than $100 h^{-1}$ Mpc. This is the first direct evidence for the cosmological principle from optical surveys (the isotropy of the CMB has for a long time been a strong point in its support). In summary, a lot of data from optical and infrared galaxies alone are currently available, and new data are being collected at a rapid rate. The observational constraints on theories of structure formation are becoming tighter. A lot of theoretical work is needed in order to allow for detailed comparisons between theory and observations. \section{Gravitational Instability} In this article we only discuss theories in which structures grow by gravitational accretion. The basic mechanism is illustrated in Fig. 20. Consider first a flat space-time background. A density perturbation with $\delta \rho > 0$ will then give rise to an excess gravitational attractive force $F$ acting on the surrounding matter. This force is proportional to $\delta \rho$, and will hence lead to exponential growth of the perturbation since $$ \delta \ddot \rho \sim F \sim \delta \rho \Rightarrow \delta \rho \sim \exp (\alpha t) \eqno\eq $$ with some constant $\alpha$. \smallskip \epsfxsize=8.2cm \epsfbox{bfig20.eps} {\baselineskip=13pt \noindent{\bf Figure 20:} Sketch of the gravitational instability mechanism. The vertical axis is the density perturbation $D$ as a function of a line in space ($x$). A small initial overdensity ($A$) will cause a gravitational acceleration $g$ towards it, which will lead to an increase in the perturbation ($B$). Note that in general underdense regions develop in addition to the growing overdense areas.} \medskip In an expanding background space-time, the acceleration is damped by the expansion. If $r (t)$ is the physcial distance of a test particle from the perturbation, then on a scale $r$ $$ \delta \ddot \rho \sim F \sim \, {\delta \rho\over{r^2 (t)}} \, , \eqno\eq $$ which results in power-law increase of $\delta \rho$. The goal of this subsection is to discuss the growth rates of inhomogeneities in more detail (see e.g. Refs. 76 and 77 for modern reviews). Because of our assumption that all perturbations start out with a small amplitude, we can linearize the equations for gravitational fluctuations. The analysis is then greatly simplified by going to Fourier space in which all modes $\delta (\underline{k})$ decouple. We expand the fractional density contrast $\delta (\underline{x})$ as follows: $$ \delta (\underline{x}) = {\delta \rho (x)\over \rho} = (2 \pi)^{-3/2} V^{1/2} \int d^3 k \> e^{i \underline{k} \cdot \underline{x}} \delta (\underline{k}) \, , \eqno\eq $$ where $V$ is a cutoff volume which disappears from all physical observables. The ``power spectrum" $P(k)$ is defined by $$ P (k) = < |\delta (k) |^2 > \, , \eqno\eq $$ where the braces denote an ensemble average (in most structure formation models, the generation of perturbations is a stochastic process, and hence observables can only be calculated by averaging over the ensemble. For observations, the braces can be viewed as an angular average). The physical measure of mass fluctuations on a length scale $\lambda$ is the r.m.s. mass fluctuation $\delta M/M (\lambda)$ on this scale. It is determined by the power spectrum in the following way. We pick a center $\underline{x}_0$ of a sphere $B_\lambda (x_0)$ of radius $\lambda$ and calculate $$ \big| {\delta M\over M} \big|^2 \, (\underline{x}_0 , \, \lambda) = \big| \int\limits_{B_\lambda (\underline{x}_0)} d^3 x \delta (\underline{x}) \, {1\over{V (B_\lambda)}} \big|^2 \, , \eqno\eq $$ where $V (B_\lambda)$ is the volume of the sphere. Inserting the Fourier decomposition (4.3) we obtain $$ \eqalign{ \big| {\delta M\over M} \big|^2 (\underline{x}_0 , \, \lambda) = & {V\over{(2 \pi)^3}} \, {1\over{V(B_\lambda )^2}} \int\limits_{B_\lambda (0)} d^3 x_1 \, \int\limits_{B_\lambda (0)} d^3 x_2 \, \int d^4 k_1 d^4 k_2 e^{i (\underline{k}_1 - \underline{k}_2) \cdot \underline{x}_0}\, \cr & e^{i \underline{k}_1 \cdot \underline{x}_1} e^{-i \underline{k}_2 \cdot \underline{x}_2} \delta (\underline{k}_1)^{\ast} \, \delta (\underline{k}_2) \, . } \eqno\eq $$ Taking the average value of this quantity over all $\underline{x}_0$ yields $$ < \left( {\delta M\over M} \right)^2 (\lambda) > = \int d^3 k W_k (\lambda) | \delta (\underline{k} |^2 \eqno\eq $$ with a window function $W_k (\lambda)$ with the following properties $$ W_k (\lambda) \cases{\simeq 1 & $k < k_\lambda = 2 \pi / \lambda$ \cr \simeq 0 & $k > k_\lambda$ .\cr} \eqno\eq $$ Therefore the r.m.s. mass perturbation on a scale $\lambda$ becomes $$ < \big| {\delta M\over M} (\lambda) \big|^2 > \sim k_{\lambda}^3 P(k_{\lambda}) \, . \eqno\eq $$ Astronomers usually assume that $P(k)$ grows as a power of $k$: $$ P (k) \sim k^n \, , \eqno\eq $$ $n$ being called the index of the power spectrum. For $n = 1$ we get the so-called Harrison-Zel'dovich scale invariant spectrum$^{78)}$. Both inflationary Universe and topological defect models of structure formation predict a roughly scale invariant spectrum. The distinguishing feature of this spectrum is that the r.m.s. mass perturbations are independent of the scale $k$ when measured at the time $t_H (k)$ when the associated wavelength is equal to the Hubble radius, i.e., when the scale ``enters" the Hubble radius. Let us derive this fact for the scales entering during the matter dominated epoch. The time $t_H (k)$ is determined by $$ k^{-1} a (t_H (k)) = t_H (k) \eqno\eq $$ which leads to $$ t_H (k) \sim k^{-3} \, . \eqno\eq $$ According to the linear theory of cosmological perturbations discussed in the following subsection, the mass fluctuations increase as $a(t)$ for $t > t_{eq}$. Hence $$ {\delta M\over M} (k, t_H (k)) = \left( {t_H (k)\over t} \right)^{2/3} \, {\delta M\over M} \, (k, t) \sim {\rm const} \, , \eqno\eq $$ since the first factor scales as $k^{-2}$ and -- using (4.15) and inserting $n=1$ -- the second as $k^2$. \section{Newtonian Theory of Cosmological Perturbations} The Newtonian theory of cosmological perturbations is an approximate analysis which is valid on wavelengths $\lambda$ much smaller than the Hubble radius $t$ and for negligible pressure $p$, i.e., $p \ll \rho$. It is based on expanding the hydrodynamical equations about a homogeneous background solution. The starting points are the continuity, Euler and Poisson equations $$ \dot \rho + \underline{\nabla} (\rho \underline{v}) = 0 \eqno\eq $$ $$ \underline{\dot v} + (\underline{v} \cdot \underline{\nabla}) \underline{v} = - \underline{\nabla} \phi - {1\over \rho} \, \underline{\nabla} p \eqno\eq $$ $$ \nabla^2 \phi = 4 \pi G \rho \eqno\eq $$ for a fluid with energy density $\rho$, pressure $p$, velocity $\underline{v}$ and Newtonian gravitational potential $\phi$, written in terms of physical coordinates $(t, \, \underline{r})$. The transition to an expanding space is made by introducing comoving coordinates $\underline{x}$ and peculiar velocity $\underline{u} = \underline{\dot {x}}$: $$ \underline{r} = a (t) \underline{x} \eqno\eq $$ $$ \underline{v} = \dot a (t) \underline{x} + a (t) \underline{u} \, . \eqno\eq $$ The first term on the right hand side of (4.24) is the expansion velocity. The perturbation equations are obtained by linearizing Equations (4.20-22) about a homogeneous background solution $\rho = \bar \rho (t) , \> p = 0$ and $\underline{u} = 0$. Using the definition $$ \delta \equiv {\delta \rho\over \rho} \, ,\eqno\eq $$ the linearization ansatz can be written $$ \rho (\underline{x}, t) = \bar \rho (t) (1 + \delta (\underline{x}, t) ) \, . \eqno\eq $$ If we consider adiabatic perturbations (no entropy density variations), then after some algebra the linearized equations become $$ \dot \delta + \nabla \cdot \underline{u} = 0 \, , \eqno\eq $$ $$ \underline{\dot {u}} + 2 {\dot a\over a} \underline{u} = - a^2 (\nabla \delta \phi + c^2_s \nabla \delta) \eqno\eq $$ and $$ \nabla^2 \delta \phi = 4 \pi G \bar \rho a^2 \delta \, , \eqno\eq $$ with the speed of sound $c_s$ given by $$ c^2_s = {\partial p\over{\partial \rho}} \, . \eqno\eq $$ The two first order equations (4.27) and (4.28) can be combined to yield a single second order differential equation for $\delta$. With the help of (4.29) this equation reads $$ \ddot \delta + 2 H \dot \delta - 4 \pi G \bar \rho \delta - {c^2_s\over a^2} \nabla^2 \delta = 0 \eqno\eq $$ which in Fourier space becomes $$ \ddot \delta_{\underline{k}} + 2 H \dot \delta_{\underline{k}} + \left( {c^2_s k^2 \over a^2} - 4 \pi G \bar \rho \right) \delta_{\underline{k}} = 0 \, . \eqno\eq $$ Here, $H(t)$ as usual denotes the expansion rate, and $\delta_{\underline{k}}$ stands for $\delta (\underline{k})$. Already a quick look at Equation (4.32) reveals the presence of a distinguished scale for cosmological perturbations, the Jeans length $$ \lambda_J = {2 \pi\over k_J} \eqno\eq $$ with $$ k^2_J = \left( {k\over a} \right)^2 = {4 \pi G \bar \rho\over c^2_s} \, . \eqno\eq $$ On length scales larger than $\lambda_J$, the spatial gradient term is negligible, and the term linear in $\delta$ in (4.32) acts like a negative mass square quadratic potential with damping due to the expansion of the Universe, in agreement with the intuitive analysis leading to (4.7) and (4.8). On length scales smaller than $\lambda_J$, however, (4.32) becomes a damped harmonic oscillator equation and perturbations on these scales decay. For $t > t_{eq}$ and for $\lambda \gg \lambda_J$, Equation (4.32) becomes $$ \ddot \delta_k + {4\over{3t}} \, \dot \delta_k - {2\over{3t^2}} \, \delta_k = 0 \eqno\eq $$ and has the general solution $$ \delta_k (t) = c_1 t^{2/3} + c_2 t^{-1} \, . \eqno\eq $$ This demonstrates that for $t > t_{eq}$ and $\lambda \gg \lambda_J$, the dominant mode of perturbations increases as $a(t)$, a result we already used in the previous subsection (see (4.19)). For $\lambda \ll \lambda_J$ and $t > t_{eq}$, Equation (4.32) becomes $$ \ddot \delta_k + 2 H \dot \delta_k + c_s^2 \left({k\over a} \right)^2 \delta_k = 0 \, , \eqno\eq $$ and has solutions corresponding to damped oscillations: $$ \delta_k (t) \sim a^{-1/2} (t) \exp \{ \pm i c_s k \int dt^\prime a (t^\prime)^{-1} \} \, . \eqno\eq $$ As an important application of the Newtonian theory of cosmological perturbations, let us compare sub-horizon scale fluctuations in a baryon-dominated Universe $(\Omega = \Omega_B = 1)$ and in a CDM-dominated Universe with $\Omega_{CDM} = 0.9$ and $\Omega = 1$. We consider scales which enter the Hubble radius at about $t_{eq}$. In the initial time interval $t_{eq} < t < t_{rec}$, the baryons are coupled to the photons. Hence, the baryonic fluid has a large pressure $p_B$ $$ p_B \simeq p_r = {1\over 3} \, \rho_r \, . \eqno\eq $$ Hence, the speed of sound is relativistic $$ c_s \simeq \left( {p_r\over \rho_m} \right)^{1/2} = {1\over{\sqrt{3}}} \, \left({\rho_r\over \rho_m} \right)^{1/2} \, . \eqno\eq $$ The value of $c_s$ slowly decreases in this time interval, attaining a value of about $1/10$ at $t_{rec}$. From (4.34) it follows that the Jeans mass $M_J$, the mass inside a sphere of radius $\lambda_J$, increases until $t_{rec}$ when it reaches its maximal value $M_J^{max}$ $$ M_J^{max} = M_J (t_{rec}) = {4 \pi\over 3} \, \lambda_J (t_{rec})^3 \bar \rho (t_{rec}) \sim 10^{17} (\Omega h^2)^{-1/2} M_{\odot} \, . \eqno\eq $$ At the time of recombination, the baryons decouple from the radiation fluid. Hence, the baryon pressure $p_B$ drops abruptly, as does the Jeans length (see (4.34)). The remaining pressure $p_B$ is determined by the temperature and thus continues to decrease as $t$ increases. It can be shown that the Jeans mass continues to decrease after $t_{rec}$, starting from a value $$ M^-_J (t_{rec}) \sim 10^6 (\Omega h^2)^{-1/2} \, M_\odot \eqno\eq $$ (where the superscript ``$-$" indicates the mass immediately after $t_{eq}$. In contrast, CDM has negligible pressure throughout the period $t > t_{eq}$ and hence experiences no Jeans damping. A CDM perturbation which enters the Hubble radius at $t_{eq}$ with amplitude $\delta_i$ has an amplitude at $t_{rec}$ given by $$ \delta^{CDM}_k (t_{rec}) \simeq \, {a (t_{rec})\over{a (t_{eq})}} \, \delta_i \, , \eqno\eq $$ whereas a perturbation with the same scale and initial amplitude in a baryon-dominated Universe is damped $$ \delta_k^{BDM} (t_{rec}) \simeq \, \left({a (t_{eq})\over{a (t_{eq})}} \right)^{-1/2} \, \delta_i \, . \eqno\eq $$ In order for the perturbations to have the same amplitude today, the initial size of the inhomogeneity must be much larger in a BDM-dominated Universe than in a CDM-dominated one: $$ \delta^{BDM}_k (t_{eq}) \simeq \left( {z (t_{eq})\over{z (t_{eq})}} \right)^{3/2} \delta_k^{CDM} \, (t_{eq}) \, . \eqno\eq $$ For $\Omega = 1$ and $h = 1/2$ the enhancement factor is about 30. In a CDM-dominated Universe the baryons experience Jeans damping, but after $t_{rec}$ the baryons quickly fall into the potential wells created by the CDM perturbations, and hence the baryon perturbations are proportional to the CDM inhomogeneities (see Fig. 21). \smallskip \epsfxsize=9cm \epsfbox{bfig21.eps} {\baselineskip=13pt \noindent{\bf Figure 21:} Comparison of the growth of fluctuations in baryon and CDM dominated Universes. The horizontal axis is time, the vertical axis is the density perturbation. The curve labelled by $a$ describes the evolution of fluctuations in a BDM Universe. The growth of CDM perturbations follows the curve $b$, and curve $c$ is a sketch of the time development of baryon perturbations in a CDM dominated Universe. To be specific, perturbations on a scale which enters the Hubble radius at $t_{eq}$ are considered.} \medskip The above considerations, coupled with information about CMB anisotropies, can be used to rule out a model with $\Omega = \Omega_B = 1$. The argument goes as follows (see Section 7). For adiabatic fluctuations, the amplitude of CMB anisotropies on an angular scale $\vartheta$ is determined by the value of $\delta \rho/\rho$ on the corresponding length scale $\lambda (\vartheta)$ at $t_{eq}$: $$ {\delta T\over T} (\vartheta) = {1\over 3} \, {{\delta \rho} \over \rho} ( \lambda (\vartheta), \, t_{eq} ) \, . \eqno\eq $$ On scales of clusters we know that (for $\Omega = 1$ and $h = 1/2$) $$ \left({{\delta \rho} \over \rho} \right)_{CDM} \, (\lambda (\vartheta), \, t_{eq}) \simeq z (t_{eq})^{-1} \simeq 10^{-4} \, , \eqno\eq $$ using the fact that today on cluster scales $\delta \rho/\rho \simeq 1$. The bounds on $\delta T/ T$ on small angular scales are $$ {\delta T\over T} << (\vartheta) 10^{-4} \, , \eqno\eq $$ consistent with the predictions for a CDM model, but inconsistent with those of a $\Omega = \Omega_B = 1$ model, according to which we would expect anisotropies of the order of $10^{-3}$. This is yet another argument in support of the existence of nonbaryonic dark matter. To conclude this subsection, let us briefly discuss Newtonian perturbations during the radiation-dominated epoch. We consider matter fluctuations with $c_s = 0$ in a smooth relativistic background. In this case, Equation (4.32) becomes $$ \ddot \delta_k + 2 H \dot \delta_k - 4 \pi G \bar \rho_m \delta_k = 0 \, , \eqno\eq $$ where $\bar \rho_m$ denotes the average matter energy density. The Hubble expansion parameter obeys $$ H^2 = {8 \pi G\over 3} (\bar \rho_m + \bar \rho_r) \, , \eqno\eq $$ with $\bar \rho_r$ the background radiation energy density. For $t < t_{eq}$, $\bar \rho_m$ is negligible in both (4.49) and (4.50), and (4.49) has the general solution $$ \delta_k (t) = c_1 \log t + c_2 \, . \eqno\eq $$ In particular, this result implies that CDM perturbations which enter the Hubble radius before $t_{eq}$ have an amplitude which grows only logarithmically in time until $t_{eq}$. This is sometimes called the Meszaros effect. \section{Relativistic Theory of Cosmological Perturbations} On scales larger than the Hubble radius $(\lambda > t)$ the Newtonian theory of cosmological perturbations obviously is inapplicable, and a general relativistic analysis is needed. On these scales, matter is essentially frozen in comoving coordinates. However, space-time fluctuations can still increase in amplitude. In principle, it is straightforward to work out the general relativistic theory of linear fluctuations$^{79)}$. We linearize the Einstein equations $$ G_{\mu\nu} = 8 \pi G T_{\mu\nu} \eqno\eq $$ (where $G_{\mu\nu}$ is the Einstein tensor associated with the space-time metric $g_{\mu\nu}$, and $T_{\mu\nu}$ is the energy-momentum tensor of matter) about an expanding FRW background $(g^{(0)}_{\mu\nu} ,\, \varphi^{(0)})$: $$ \eqalign{ g_{\mu\nu} (\underline{x}, t) & = g^{(0)}_{\mu\nu} (t) + h_{\mu\nu} (\underline{x}, t) \cr \varphi (\underline{x}, t) & = \varphi^{(0)} (t) + \delta \varphi (\underline{x}, t) \, \cr} \eqno\eq $$ and pick out the terms linear in $h_{\mu\nu}$ and $\delta \varphi$ to obtain $$ \delta G_{\mu\nu} \> = \> 8 \pi G \delta T_{\mu\nu} \, . \eqno\eq $$ In the above, $h_{\mu\nu}$ is the perturbation in the metric and $\delta \varphi$ is the fluctuation of the matter field $\varphi$. We have denoted all matter fields collectively by $\varphi$. In practice, there are many complications which make this analysis highly nontrivial. The first problem is ``gauge invariance"$^{80)}$ Imagine starting with a homogeneous FRW cosmology and introducing new coordinates which mix $\underline{x}$ and $t$. In terms of the new coordinates, the metric now looks inhomogeneous. The inhomogeneous piece of the metric, however, must be a pure coordinate (or "gauge") artefact. Thus, when analyzing relativistic perturbations, care must be taken to factor out effects due to coordinate transformations. \smallskip \epsfxsize=9.5cm \epsfbox{bfig22.eps} {\baselineskip=13pt \noindent{\bf Figure 22:} Sketch of how two choices of the mapping from the background space-time manifold ${\cal M}_0$ to the physical manifold ${\cal M}$ induce two different coordinate systems on ${\cal M}$.} \medskip The issue of gauge dependence is illustrated in Fig. 22. A coordinate system on the physical inhomogeneous space-time manifold ${\cal M}$ can be viewed as a mapping ${\cal D}$ of an unperturbed space-time ${\cal M}_0$ into ${\cal M}$. A physical quantity $Q$ is a geometrical function defined on ${\cal M}$. There is a corresponding physical quantity $^{(0)}Q$ defined on ${\cal M}_0$. In the coordinate system given by ${\cal D}$, the perturbation $\delta Q$ of $Q$ at the space-time point $p \, \epsilon \, {\cal M}$ is $$ \delta Q (p) = Q (p) - \,^{(0)}Q \, (D^{-1} (p) ) \, . \eqno\eq $$ However, in a second coordinate system $\tilde {\cal D}$ the perturbation is given by $$ \delta \tilde Q (p) = Q (p) - \,^{(0)}Q (\tilde {\cal D}^{-1} (p) ) \, . \eqno\eq $$ The difference $$ \Delta Q (p) = \delta Q (p) - \delta \tilde Q (p) \eqno\eq $$ is obviously a gauge artefact and carries no physical meaning. There are various methods of dealing with gauge artefacts. The simplest and most physical approach is to focus on gauge invariant variables, i.e., combinations of the metric and matter perturbations which are invariant under linear coordinate transformations. The gauge invariant theory of cosmological perturbations is in principle straightforward, although technically rather tedious. In the following I will summarize the main steps and refer the reader to Ref. 11 for the details and further references (see also Ref. 81 for a pedagogical introduction and Refs. 82-87 for other approaches). We consider perturbations about a spatially flat Friedmann-Robertson-Walker metric $$ ds^2 = a^2 (\eta) (d\eta^2 - d \underline{x}^2) \eqno\eq $$ where $\eta$ is conformal time (related to cosmic time $t$ by $a(\eta) d \eta = dt$). A scalar metric perturbation (see Ref. 88 for a precise definition) can be written in terms of four free functions of space and time: $$ \delta g_{\mu\nu} = a^2 (\eta) \pmatrix{2 \phi & -B_{,i} \cr -B_{,i} & 2 (\psi \delta_{ij} + E_{,ij} \cr} \, . \eqno\eq $$ Scalar metric perturbations are the only perturbations which couple to energy density and pressure. The next step is to consider infinitesimal coordinate transformations $$ x^{\mu^\prime} = x^\mu + \xi^\mu \eqno\eq $$ which preserve the scalar nature of $\delta g_{\mu\nu}$ and to calculate the induced transformations of $\phi, \psi, B$ and $E$. Then we find invariant combinations to linear order. (Note that there are in general no combinations which are invariant to all orders$^{89)}$.) After some algebra, it follows that $$ \eqalign{ \Phi & = \phi + a^{-1} [(B - E^\prime) a]^\prime \cr \Psi & = \psi - {a^\prime\over a} \, (B - E^\prime) \cr} \eqno\eq $$ are two invariant combinations. In the above, a prime denotes differentiation with respect to $\eta$. There are various methods to derive the equations of motion for gauge invariant variables. Perhaps the simplest way$^{11)}$ is to consider the linearized Einstein equations (4.54) and to write them out in the longitudinal gauge defined by $$ B = E = 0 \eqno\eq $$ and in which $\Phi = \phi$ and $\Psi = \psi$, to directly obtain gauge invariant equations. For several types of matter, in particular for scalar field matter, the perturbation of $T_{\mu \nu}$ has the special property $$ \delta T^i_j \sim \delta^i_j \eqno\eq $$ which imples $\Phi = \Psi$. Hence, the scalar-type cosmological perturbations can in this case be described by a single gauge invariant variable. The equation of motion takes the form$^{90, 9, 10)}$ $$ \dot \xi = O \left({k\over{aH}} \right)^2 H \xi \eqno\eq $$ where $$ \xi = {2\over 3} \, {H^{-1} \dot \Phi + \Phi\over{1 + w}} + \Phi \, . \eqno\eq $$ The variable $w = p/ \rho$ (with $p$ and $\rho$ background pressure and energy density respectively) is a measure of the background equation of state. In particular, on scales larger than the Hubble radius, the right hand side of (4.64) is negligible, and hence $\xi$ is constant. The result that $\dot \xi = 0$ is a very powerful one. Let us first imagine that the equation of state of matter is constant, {\it i.e.}, $w = {\rm const}$. In this case, $\dot \xi = 0$ implies $$ \Phi (t) = {\rm const} \, , \eqno\eq $$ {\it i.e.}, this gauge invariant measure of perturbations remains constant outside the Hubble radius. Next, consider the evolution of $\Phi$ during a phase transition from an initial phase with $w = w_i$ to a phase with $w = w_f$. Long before and after the transition, $\Phi$ is constant because of (4.66), and hence $\dot \xi = 0$ becomes $$ {\Phi\over{1 + w}} + \Phi = {\rm const} \, , \eqno\eq $$ In order to make contact with matter perturbations and Newtonian intuition, it is important to remark that, as a consequence of the Einstein constraint equations, at Hubble radius crossing $\Phi$ is a measure of the fractional density fluctuations: $$ \Phi (k, t_H (k) ) \sim {\delta \rho\over \rho} \, ( k , \, t_H (k) ) \, . \eqno\eq $$ (Note that the latter quantity is approximately gauge invariant on scales smaller than the Hubble radius). \chapter{Inflationary Universe Scenarios} \section{Preliminaries} Cosmological inflation$^{30)}$ is a period in time during which the Universe is expanding exponentially, {\it i.e.}, $$ a (t) = e^{tH} \eqno\eq $$ with constant Hubble expansion rate $H$. From the FRW equations (2.7) and (2.9) it follows that the condition for inflation (in the context of Einstein gravity in a spatially flat Universe) is an equation of state for matter with $$ p = - \rho \, , \eqno\eq $$ which neccessitates abandoning a description of matter in terms of an ideal gas. As was indicated in Section 3.1, it is possible to achieve inflation if matter is described in terms of scalar fields, provided that at some period $$ \eqalign{ \dot \varphi^2 \ll V (\varphi) \cr (\nabla \varphi)^2 \ll V (\varphi) }\eqno\eq $$ (see Eqs. (3.9) which give the equation of state for scalar field matter). \smallskip \epsfxsize=9cm \epsfbox{bfig23.eps} {\baselineskip=13pt \noindent{\bf Figure 23:} A sketch of two potentials which can give rise to inflation.} \medskip Two examples which can give inflation are shown in Fig. 23. In (a), inflation occurs at the stable fixed point $\varphi (\underline{x}, t_i) = 0 = \dot \varphi (\underline{x}, t_i)$. However, this model is ruled out by observation: the inflationary phase has no ending. $V(0)$ acts as a permanent nonvanishing cosmological constant. In (b), a finite period of inflation can arise if $\varphi (\underline{x})$ is trapped at the local minimum $\varphi = 0$ with $\dot \varphi (\underline{x}) = 0$. However, in this case $\varphi (\underline{x})$ can make a sudden transition at some time $t_R > t_i$ through the potential barrier and move to $\varphi (\underline{x}) = a$. Thus, for $t_i < t < t_R$ the Universe expands exponentially, whereas for $t > t_R$ the contribution of $\varphi$ to the expansion of the Universe vanishes and we get the usual FRW cosmology. There are three obvious questions: why does the field start out at $\varphi = 0$, how does the transition occur and why should the scalar field have $V(\varphi) = 0$ at the global minimum? In the following section the first two questions will be addressed. The third question is part of the cosmological constant problem for which there is as yet no convincing explanation. Before studying the dynamics of the phase transition, we need to digress and discuss finite temperature effects. \section{Finite Temperature Field Theory} \par The evolution of particles in vacuum and in a thermal bath are very different. Similarly, the evolution of fields changes when coupled to a thermal bath. Under certain conditions, the changes may be absorbed in a temperature dependent potential, the finite temperature effective potential. Here, a heuristic derivation of this potential will be given. The reader is referred to Ref. 8 or to the original articles$^{91)}$ for the actual derivation. We assume that the scalar field $\varphi (\undertext{x},t)$ is coupled to a thermal bath which is represented by a second scalar field $\psi (\undertext{x}, t)$ in thermal equilibrium. The Lagrangian for $\varphi$ is $$ {\cal L} = {1\over 2} \partial_\mu \varphi \partial^\mu \varphi \, - \, V(\varphi) \, - \, {1\over 2} \hat \lambda \varphi^2 \psi^2\, , \eqno\eq $$ where $\hat \lambda$ is a coupling constant. The action from which the equations of motion are derived is $$ S \, = \, \int d^4 x \, \sqrt{-g} {\cal L}\eqno\eq $$ where $g$ is the determinant of the metric (2.2). The resulting equation of motion for $\varphi (\undertext{x},t)$ is $$ \ddot \varphi + 3 H \dot \varphi \, - a^{-2} \bigtriangledown^2 \varphi \, = \,- V^\prime (\varphi) - \hat \lambda \psi^2 \varphi \, . \eqno\eq $$ If $\psi$ is in thermal equilibrium, we may replace $\psi^2$ by its thermal expectation value $<\psi^2>_T$. Now, $$ <\psi^2>_T \sim T^2\eqno\eq $$ which can be seen as follows: in thermal equilibrium, the energy density of $\psi$ equals that of one degree of freedom in the thermal bath. In particular, the potential energy density $V (\psi)$ of $\psi$ is of that order of magnitude. Let $$ V (\psi) \, = \, \lambda_\psi \psi^4\eqno\eq $$ with a coupling constant $\lambda_\psi$ which we take to be of the order 1 (if $\lambda_\psi$ is too small, $\psi$ will not be in thermal equilibrium). Since the thermal energy density is proportional to $T^4$, (5.7) follows. (5.6) can be rewritten as $$ \ddot \psi + 3H \dot \varphi \, - \, a^{-2} \bigtriangledown^2 \varphi \, = \, - V_T^\prime (\varphi),\eqno\eq $$ where $$ V_T (\varphi) \, = \, V (\varphi) \, + \, {1\over 2} \hat \lambda T^2 \varphi^2\eqno\eq $$ is called the finite temperature effective potential. Note that in (5.10), $\hat \lambda$ has been rescaled to absorb the constant of proportionality in (5.7). \par These considerations will now be applied to Example A, a scalar field model with potential $$ V (\varphi) \, = \, {1\over 4} \lambda (\varphi^2 - \sigma^2)^2\eqno\eq $$ ($\sigma$ is called the scale of symmetry breaking). The finite temperature effective potential becomes (see Fig. 24) $$ V_T (\varphi) \, = \, {1\over 4} \lambda \varphi^4 - {1\over 2} \, \left(\lambda \sigma^2 - \hat \lambda T^2\right) \varphi^2 + \, {1\over 4} \, \lambda \sigma^4 \, . \eqno\eq $$ For very high temperatures, the effective mass term is positive and hence the energetically favorable state is $<\varphi> = 0$. For very low temperatures, on the other hand, the mass term has a negative sign which leads to spontaneous symmetry breaking. The temperature at which the mass term vanishes defines the critical temperature $T_c$ $$ T_c \, = \, \hat \lambda^{-1/2} \lambda^{1/2} \sigma\,.\eqno\eq $$ \smallskip \epsfxsize=8cm \epsfbox{bfig24.eps} {\baselineskip=13pt \noindent {\bf Figure 24:} The finite temperature effective potential for Example A.} \medskip As Example B, consider a theory with potential $$ V (\varphi) = \, {1\over 4} \varphi^4 - {1\over 3} \, (a + b) \varphi^3 + {1\over 2} ab \varphi^2\eqno\eq $$ with ${1\over 2} a > b > 0$. The finite temperature effective potential is obtained by adding ${1\over 2} \hat \lambda T^2 \varphi^2$ to the right hand side of (5.14). ${ V_T} (\varphi)$ is sketched in Fig. 25 for various values of $ T$. The critical temperature $T_c$ is defined as the temperature when the two minima of ${V_T}(\varphi)$ become degenerate. \smallskip \epsfxsize=10cm \epsfbox{bfig25.eps} {\baselineskip=13pt \noindent {\bf Figure 25:} The finite temperature effective potential for Example B.} \medskip It is important to note that the use of finite temperature effective potential methods is only legitimate if the system is in thermal equilibrium. This point was stressed in Refs. 92 and 93, although the conclusion should be obvious from the derivation given above. To be more precise, we require the $\psi$ field to be in thermal equilibrium and the coupling constant $\hat \lambda$ of (5.4) which mediates the energy exchange between the $\varphi$ and $\psi$ fields to be large. However, as shown in Chapter 4, in inflationary Universe models, the observational constraints stemming from the amplitude of the primordial energy density fluctuation spectrum force the self coupling constant $\lambda$ of $\varphi$ to be extremely small. Since at one loop order, the interaction term ${1\over 2} \hat \lambda \varphi^2 \psi^2$ induces contributions to $\lambda$, it is unnatural to have $\lambda$ very small and $\hat \lambda$ unsuppressed. Hence, in many inflationary Universe models - in particular in new inflation$^{94)}$ and in chaotic inflation$^{92)}$ - finite temperature effective potential methods are inapplicable. \section{Phase Transitions} \par The temperature dependence of the finite temperature effective potential in quantum field theory leads to phase transitions in the very early Universe. These transitions are either first or second order. \par Example $A$ of the previous section provides a model in which the transition - at least according to the above mean field analysis - is second order (see Fig. 24). For $T \gg T_c$, the expectation value of the scalar field $\varphi$ vanishes at all points $\undertext{x}$ in space: $$ < \varphi (\undertext{x}) > = 0 \, . \eqno\eq $$ For $T < T_c$, this value of $< \varphi (\undertext{x}) >$ becomes unstable and $< \varphi (\undertext{x})>$ evolves smoothly in time to a new value $\pm \sigma$. The direction is determined by thermal and quantum fluctuations and is therefore not uniform in space. There will be domains of average radius $\xi (t)$ in which $< \varphi (\undertext{x}) >$ is coherent. By causality, the coherence length is bounded from above by the horizon. However, typical values of $\xi (t)$ are proportional to $\lambda^{-1} \sigma^{-1}$ if $\varphi$ was in thermal equilibrium before the phase transition. \par In condensed matter physics, a transition of the above type is said to proceed by spinodal decomposition$^{95)}$, triggered by a rapid quench. \par In Example B of the previous section, (see Fig. 25) the phase transition is first order. For $T > T_c$, the expectation value $<\varphi (x) >$ is approximately $0$, the minimum of the high temperature effective potential. Provided the zero temperature potential has a sufficiently high barrier separating the metastable state $\varphi = 0$ from the global minimum (compared to the energy density in thermal fluctuations at $T = T_c$), then $\varphi (\undertext{x})$ will remain trapped at $\varphi = 0$ also for $T < T_c$. In the notation of Ref. 96, the field $\varphi$ is trapped in the false vacuum. After some time (determined again by the potential barrier), the false vacuum will decay by quantum tunnelling. \par Tunnelling in quantum field theory was discussed in Refs. 96-99 (for reviews see e.g., Refs. 100 and 8). The transition proceeds by bubble nucleation. There is a probability per unit time and volume that at a point $\undertext{x}$ in space a bubble of ``true vacuum" $\varphi (\undertext{x}) = a$ will nucleate. The nucleation radius is microscopical. As long as the potential barrier is large, the bubble radius will increase with the speed of light after nucleation. Thus, a bubble of $\varphi = a$ expands in a surrounding ``sea" of false vacuum $\varphi = 0$. \par To conclude, let us stress the most important differences between the two types of phase transitions discussed above. In a second order transition, the dynamics is determined mainly by classical physics. The transition occurs homogeneously in space (apart from the phase boundaries which -- as discussed below -- become topological defects), and $< \varphi (x) >$ evolves continuously in time. In first order transitions, quantum mechanics is essential. The process is extremely inhomogeneous, and $< \varphi (x) >$ is discontinuous as a function of time. As we shall see in the following sections, the above two types of transitions are the basis of various classes of inflationary Universe models. \section{Models of Inflation} At this stage we have established the formalism to be able to discuss models of inflation. I will focus on ``old inflation," ``new inflation"" and ``chaotic inflation." There are many other attempts at producing an inflationary scenario, but there is as of now no convincing realization. \subsection{Old Inflation} \par The old inflationary Universe model$^{30, 101)}$ is based on a scalar field theory which undergoes a first order phase transition. As a toy model, consider a scalar field theory with the potential $V (\varphi)$ of Example B (see Fig. 25). Note that this potential is fairly general apart from the requirement that $V (a) = 0$, where $\varphi = a$ is the global minimum of $V (\varphi)$. This condition is required to avoid a large cosmological constant today (no inflationary Universe model manages to circumvent or solve the cosmological constant problem). \par For fairly general initial conditions, $\varphi (x)$ is trapped in the metastable state $\varphi = 0$ as the Universe cools below the critical temperature $T_c$. As the Universe expands further, all contributions to the energy-momentum tensor $T_{\mu \nu}$ except for the contribution $$ T_{\mu \nu} \sim V(\varphi) g_{\mu \nu} \eqno\eq $$ redshift. Hence, the equation of state approaches $p = - \rho$, and inflation sets in. Inflation lasts until the false vacuum decays. During inflation, the Hubble constant is given by $$ H^2 = {8 \pi G\over 3} \, V (0) \, . \eqno\eq $$ \smallskip \epsfxsize=6cm \epsfbox{bfig26.eps} {\baselineskip=13pt \noindent {\bf Figure 26:} A sketch of the spatially inhomogeneous distribution of $\varphi$ in Old Inflation.} \medskip After a period $\Gamma^{-1}$, where $\Gamma$ is the tunnelling decay rate, bubbles of $\varphi = a$ begin to nucleate in a sea of false vacuum $\varphi = 0$. For a sketch of the resulting inhomogeneous distribution of $\varphi (x)$ see Fig. 26. Note that inflation stops after bubble nucleation. \par The time evolution in old inflation is summarized in Fig. 27. We denote the beginning of inflation by $t_i$ (here $t_i \simeq t_c$), the end by $t_R$ (here $t_R \simeq t_c + \Gamma^{-1}$). \smallskip \epsfxsize=9cm \epsfbox{bfig8.eps} {\baselineskip=13pt \noindent {\bf Figure 27:} Phases in the old and new inflationary Universe.} \medskip It was immediately realized that old inflation has a serious ``graceful exit" problem$^{102)}$. The bubbles nucleate after inflation with radius $r \ll 2t_R$ and would today be much smaller than our apparent horizon. Thus, unless bubbles percolate, the model predicts extremely large inhomogeneities inside the Hubble radius, in contradiction with the observed isotropy of the microwave background radiation. \par For bubbles to percolate, a sufficiently large number must be produced so that they collide and homogenize over a scale larger than the present Hubble radius. However, with exponential expansion, the volume between bubbles expands exponentially whereas the volume inside bubbles expands only with a low power. This prevents percolation. \subsection{New Inflation} Because of the graceful exit problem, old inflation never was considered to be a viable cosmological model. However, soon after the seminal paper by Guth, Linde and independently Albrecht and Steinhardt put forwards a modified scenario, the New Inflationary Universe$^{94)}$ (see also Ref. 103). \smallskip \epsfxsize=6cm \epsfbox{bfig28.eps} {\baselineskip=13pt \noindent {\bf Figure 28:} A sketch of the spatial distribution of $\varphi$ in New Inflation after the transition. The symbols $+$ and $-$ indicate regions where $\varphi = + \sigma$ and $\varphi = - \sigma$ respectively.} \medskip The starting point is a scalar field theory with a double well potential which undergoes a second order phase transition (Fig. 24). $V(\varphi)$ is symmetric and $\varphi = 0$ is a local maximum of the zero temperature potential. Once again, it was argued that finite temperature effects confine $\varphi(\undertext{x})$ to values near $\varphi = 0$ at temperatures $T \ge T_c$. For $T < T_c$, thermal fluctuations trigger the instability of $\varphi (\undertext{x}) = 0$ and $\varphi (\undertext{x})$ evolves towards $\varphi = \pm \sigma$ by the classical equation of motion $$ \ddot \varphi + 3 H \dot \varphi - a^{-2} \bigtriangledown^2 \varphi = - V^\prime (\varphi)\, . \eqno\eq $$ \par The transition proceeds by spinodal decomposition (see Fig. 28) and hence $\varphi(\undertext{x})$ will be homogeneous within a correlation length. The analysis will be confined to such a small region. Hence, in Eq. (5.18) we can neglect the spatial gradient terms. Then, from (3.9) we can read off the induced equation of state. The condition for inflation is $$ \dot \varphi^2 \ll V (\varphi)\, ,\eqno\eq $$ \ie~ slow rolling. \par Often, the ``slow rolling" approximation is made to find solutions of (5.18). This consists of dropping the $\ddot \varphi$ term. In this case, (5.18) becomes $$ 3 H \dot \varphi \, = - \, V^\prime (\varphi)\, . \eqno\eq $$ As an example, consider a potential which for $|\varphi| < \sigma$ has the following expansion near $\varphi = 0$ $$ V (\varphi) = V_0 - {1\over 2} m^2 \varphi^2\, . \eqno\eq $$ With the above $V (\phi)$, (5.20) has the solution $$ \varphi (t) \, = \, \varphi (0) \exp \left({m^2\over{3H}} t \right)\eqno\eq $$ (taking $H = \rm const$ which is a good approximation). Thus, provided $m \ll \sqrt{3} H \, , \, \ddot \varphi$ is indeed smaller than the other terms in (5.18) and the slow rolling approximation seems to be satisfied. \par However, the above conclusion is premature$^{104)}$. Equation (5.18) has a second solution. For $m > H$ the solution is $$ \varphi (t) \simeq \varphi (0) e^{mt}\eqno\eq $$ and dominates over the previous one. This example shows that the slow rolling approximation must be used with caution. Here, however, the conclusion remains that provided $m \ll H$, then the model produces enough inflation to solve the cosmological problems. \par There is no graceful exit problem in the new inflationary Universe. Since the spinodal decomposition domains are established before the onset of inflation, any boundary walls will be inflated outside the present Hubble radius. \par The condition $m^2 \ll 3 H^2$ which must be imposed in order to obtain inflation, is a fine tuning of the particle physics model -- the first sign of problems with this scenario. Consider \eg~ the model (5.11). By expanding $V (\varphi)$ about $\varphi = 0$ we can determine both $H$ and $m$ in terms of $\lambda$ and $\sigma$. In order that $m^2 < 3 H^2$ be satisfied we need $$ \sigma > \, \left( {1 \over{6 \pi}}\right)^{1/2} m_{pl} \, , \eqno\eq $$ which is certainly an unnatural constraint for models motivated by particle physics. \par Let us, for the moment, return to the general features of the new inflationary Universe scenario. At the time $t_c$ of the phase transition, $\varphi (t)$ will start to move from near $\varphi = 0$ towards either $\pm \sigma$ as described by the classical equation of motion, \ie~ (5.22). At or soon after $t_c$, the energy-momentum tensor of the Universe will start to be dominated by $V(\varphi)$, and inflation will commence. $t_i$ shall denote the time of the onset of inflation. Eventually, $\phi (t)$ will reach large values for which (5.21) is no longer a good approximation to $V (\varphi)$ and for which nonlinear effects become important. The time at which this occurs is $t_B$. For $t > t_B \, , \, \varphi (t)$ rapidly accelerates, reaches $\pm \sigma$, overshoots and starts oscillating about the global minimum of $V (\varphi)$. The amplitude of this oscillation is damped by the expansion of the Universe and (predominantly) by the coupling of $\varphi$ to other fields. At time $t_R$, the energy in $\varphi$ drops below the energy of the thermal bath of particles produced during the period of oscillation. \par The evolution of $\varphi (t)$ is sketched in Fig. 29. The time period between $t_B$ and $t_R$ is called the reheating period and is usually short compared to the Hubble expansion time. The time evolution of the temperature $T$ of the thermal radiation bath is also shown in Fig. 29. \smallskip \epsfxsize=10cm \epsfbox{bfig29.eps} {\baselineskip=13pt \noindent {\bf Figure 29:} Evolution of $\varphi (t)$ and $T (t)$ in the new inflationary Universe.} \medskip Reheating in inflationary Universe models has been considered in Refs. 105-108. One way to view the process is as follows$^{107, 108)}$. Consider a second scalar field $\psi$ coupled to $\varphi$ via the interaction Lagrangian $$ {\cal L}_I = {1\over 2} g \varphi^2 \psi^2\, . \eqno\eq $$ Then, an oscillating $\varphi (t)$ will act as a time dependent mass with periodic variations in the equation of motion for $\psi$ $$ \ddot \psi_k + 3 H \dot \psi_k + \left(m_\psi^2 + k^2 a^{-2} (t) + g \varphi^2 (t) \right) \psi_k = 0\eqno\eq $$ where we have neglected nonlinear terms and expanded $\psi$ into Fourier modes $\psi_k$. If the expansion of the Universe can be neglected and for periodic $\varphi^2 (t)$, the above is the well known Mathieu equation$^{109)}$ whose solutions have instabilities for certain values of $k$. These instabilities correspond to the production of $\psi$ particles with well determined momenta$^{107, 108)}$. These particles eventually equilibrate and regenerate a thermal bath. \par For $t > t_R$, the Universe is again radiation dominated. Hence, the stages of the new inflationary Universe are the same as for old inflation (Fig. 27). There is a useful order of magnitude relation between the scale of symmetry breaking $\sigma$ and $H$. From $$ H^2 = \, {8 \pi G\over 3} \, V(0)\eqno\eq $$ and from the form of the potential (see (5.11)) it follows that $$ \left({H\over \sigma}\right) \sim \, \lambda^{1/2} \, \left({\sigma\over{m_{pl}}}\right)\, .\eqno\eq $$ In particular, for $\sigma \sim 10^{15}$GeV (typical scale of grand unification) and $\lambda \sim 1$ we obtain $H \sim 10^{11}$GeV. \par The new inflationary Universe model -- although it was for a long time presented as a viable model -- suffers from severe fine tuning and initial condition problems. In (5.24) we encountered the first of these problems: in order to obtain enough inflation, the potential must be fairly flat near $\varphi = 0$. A more severe problem will be derived in Section 5.5: Inflationary Universe models generate energy density perturbations. The steeper the potential, the larger the density perturbations. For a potential which near $\varphi = 0$ or $\varphi = H$ has the following expansion $$ V (\varphi) = V (0) - \lambda \, \varphi^4\, ,\eqno\eq $$ the density perturbations conflict with observations unless (see later) $$ \lambda < 10^{-12}\, .\eqno\eq $$ \par This in itself is an unexplained small number problem. However, even if we were willing to accept this we would run into initial condition problems$^{92, 93)}$. For the new inflationary Universe to proceed in the way outlined above, it is essential that the field $\varphi$ be in thermal equilibrium with other fields. This implies that the constant $g$ coupling $\varphi$ to other fields should not be too small. However, a coupling term of the form (5.25) induces one loop quantum corrections to the self coupling constant $\lambda$ of the order $g^2$. Hence, the constraint $\lambda < 10^{-12}$ implies a constraint $g < 10^{-6}$. Thus, $\varphi$ will not be in thermal equilibrium at $t_c$, and hence there will be no thermal forces which localize $\varphi$ close to $\varphi = 0$. \par Note that the above problem is not an artifact of using quartic potentials such as (5.29). Similar constraints would arise in other (\eg~ quadratic) models. However, (5.29) was long considered to be the prototypical shape of $V (\varphi)$ for small values of $\varphi$ since it is the shape which arises in Coleman-Weinberg$^{110)}$ models. \par In the absence of thermal forces which constrain $\varphi$ to start close to $\varphi = 0$, the only constraints on $\varphi - \,$ at least using classical physics alone -- come from energetic considerations. Obviously, it is unnatural to assume that at the initial time $t_i$ the energy density in $\varphi$ exceeds the energy density of one degree of freedom of the thermal bath at time $t_i$ (temperature $T_i$). This implies $$ \eqalign{V (\varphi (\undertext{x}, t_i)) & < {\pi^2\over{30}} T_i^4\cr | \bigtriangledown \varphi (\undertext{x}, t_i) |^2 & < {\pi^2\over{30}} T_i^4 a^2 (t_i)\cr | \dot \varphi^2 (\undertext{x}, t_i) |^2 &< {\pi^2\over{30}} T_i^4}\eqno\eq $$ In particular, for the double well potential of (5.11), (5.31) implies that $\varphi (\undertext{x}, t_i)$ can be of the order $$ \varphi (\undertext{x}, t_i) \sim \lambda^{-1/4} T_i\eqno\eq $$ which for $T_i > \sigma$ is much larger than $\sigma$. In a weakly coupled model, the only natural time to impose initial conditions on $\varphi (\undertext{x})$ is the Planck time, \ie~ $T_i \sim m_{pl}$. Hence, the initial conditions allow and in fact suggest $$ \varphi (\undertext{x}, t_i) \sim \lambda^{-1/4} m_{pl} \, \gg \, m_{pl} \quad \rm{for} \> T_i \sim T_{pl},\eqno\eq $$ These observations lead to the chaotic inflation scenario, the only of the original inflationary Universe models which can still be considered as a viable scenario today. \subsection{Chaotic Inflation} \par Chaotic inflation$^{92)}$ is based on the observation that for weakly coupled scalar fields, initial conditions which follow from classical considerations alone lead to very large values of $\varphi (\undertext{x})$ (see (5.32)). \par Consider a region in space where at the initial time $\varphi (\undertext{x})$ is very large, homogeneous (we will make these assumptions quantitative below) and static. In this case, the energy-momentum tensor will be immediately dominated by the large potential energy term and induce an equation of state $p \simeq - \rho$ which leads to inflation. Due to the large Hubble damping term in the scalar field equation of motion, $\varphi (\undertext{x})$ will only roll very slowly towards $\varphi = 0$. The kinetic energy contribution to $T_{\mu \nu}$ will remain small, the spatial gradient contribution will be exponentially suppressed due to the expansion of the Universe, and thus inflation persists. This is a brief survey of the chaotic inflation scenario. Note that in contrast to old and new inflation, no initial thermal bath is required. Note also that the precise form of $V(\varphi)$ is irrelevant to the mechanism. In particular, $V(\varphi)$ need not be a double well potential. This is a significant advantage, since for scalar fields other than Higgs fields used for spontaneous symmetry breaking, there is no particle physics motivation for assuming a double well potential, and since the inflaton (the field which gives rise to inflation) cannot be a conventional Higgs field due to the severe fine tuning constraints. \par Let us consider the chaotic inflation scenario in more detail. For simplicity, take the potential $$ V (\varphi) = \, {1\over 2} \, m^2 \varphi^2\eqno\eq $$ and consider a region in space in which $\varphi (\undertext{x}, t_i)$ is sufficiently homogeneous. To be specific, we require $$ {1\over 2} \, a^{-2} (t_i) | \bigtriangledown \varphi \, (\undertext{x}, t_i)|^2 \, \ll \, V\left(\varphi (\undertext{x}, t_i)\right)\eqno\eq $$ over a region of size $d_i$ $$ d_i \ge 3 H^{-1} (t_i)\, .\eqno\eq $$ We also require that the kinetic energy be negligible at the initial time $t_i$, $$ \dot \varphi (\undertext{x}, t_i)^2 \, \ll \, V (\varphi \left(\undertext{x}, t_i) \right)\, ,\eqno\eq $$ although this assumption can be relaxed without changing the results$^{111)}$. {}From (5.35) and (5.37) it follows that at $t_i$ the equation of state is inflationary, \ie~ $p (t_i) \simeq - \rho (t_i)$. Condition (5.36) ensures that no large inhomogeneities can propagate from outside to the center of the region under consideration. With these approximations, the equation of motion for $\varphi$ becomes $$ \ddot \varphi + 3 H \dot \varphi = - m^2 \varphi\eqno\eq $$ with $$ H = \left({4 \pi\over 3}\right)^{1/2} \, {m\over{m_{pl}}} \, \varphi\, . \eqno\eq $$ \par Since we expect $\varphi (\undertext{x}, t)$ to be changing slowly, we make the slow rolling approximation $$ 3 H \dot \varphi = - m^2 \varphi\eqno\eq $$ which gives $$ \dot \varphi = - \left({1\over{12 \pi}}\right)^{1/2} m \, m_{pl}\eqno\eq $$ and shows that the approximation is self consistent. In order to get inflation, we require $$ {1\over 2} \, \dot \varphi^2 < {1\over 2} m^2 \, \varphi^2\eqno\eq $$ which (by (5.41)) is satisfied if $$ \varphi > \left({1\over{12}}\right)^{1/2} m_{pl}\, . \eqno\eq $$ In order to obtain a period $\tau > 50 H^{-1} (t_i)$ of inflation, a slightly stronger condition is needed: $$ \varphi > 3 m_{pl}\, . \eqno\eq $$ \par With chaotic inflation, the initial hope that grand unified theories could provide the answer to the homogeneity and flatness problems has been abandoned. The inflaton is introduced as a new scalar field (with no particular particle physics role) which is very weakly coupled to itself and to other fields (see Ref. 112 for an attempt to couple the inflaton to non-grand unified particle physics). In supergravity and in superstring inspired models there are scalar fields which are candidates to be the inflaton. I refer the reader to Refs. 113-115 for a discussion of this issue. However, even in such models the time when the inflaton $\varphi$ decouples from the rest of physics is the Planck time $t_i = t_{pl}$. Thus, the chaotic inflation scenario is often called primordial inflation. \par Chaotic inflation is a much more radical departure from standard cosmology than old and new inflation. In the latter, the inflationary phase can be viewed as a short phase of exponential expansion bounded at both ends by phases of radiation domination. In chaotic inflation, a piece of the Universe emerges with an inflationary equation of state immediately after the quantum gravity epoch. The chaotic inflationary Universe scenario has been developed in great detail (see {\it e.g.}, Ref. 116 for a recent review). One important addition is the inclusion of stochastic noise$^{117)}$ in the equation of motion for $\varphi$ in order to take into account the effects of quantum fluctuations. It can in fact be shown that for sufficiently large values of $| \varphi |$, the stochastic force terms are more important than the classical relaxation force $V^\prime (\varphi)$. There is equal probability for the quantum fluctuations to lead to an increase or decrease of $| \varphi |$. Hence, in a substantial fraction of comoving volume, the field $\varphi$ will climb up the potential. This leads to the conclusion that chaotic inflation is eternal. At all times, a large fraction of the physical space will be inflating. Another consequence of including stochastic terms is that on large scales (much larger than the present Hubble radius), the Universe will look extremely inhomogeneous. \subsection{General Comments} Old, new and choatic inflation are all based on the use of new fundamental scalar fields which cannot be the Higgs field of an ordinary gauge theory. Instead of introducing new physics via scalar fields -- an approach which makes the cosmological constant problem more severe -- it is possible to look for realizations of inflation based on some alternative new physics which do not invoke fundamental scalar fields. One possibility is to consider modifications of Einstein gravity which can lead to inflation. In fact, the first model of inflation$^{39)}$ was based on considering an action for gravity of the form $$ S = \int d^4 x \sqrt{-g} \, (R + \varepsilon R^2) \, , \eqno\eq $$ where $R$ is the Ricci scalar of the metric $g_{\mu\nu}$. As shown first by Whitt$^{118)}$, the equations of motion resulting from this action are the same as those from Einstein gravity in the presence of a scalar field with a special potential which allows for chaotic inflation. The relationship is obtained via a conformal transformation. Since perturbative quantum gravity calculations and considerations based on quantum field theory in curved space-time all point to the presence of higher derivative terms in the action for gravity, it is not unlikely that successful inflation will come from the gravity sector. A recently proposed theory$^{41)}$, in which the curvature is bounded for all solutions, predicts that our Universe will have started out in an inflationary period. Another interesting possibility is that inflation is a result of new physics associated with a unified theory of all forces such as superstring theory. Interesting speculations along these lines have recently been made in Ref. 119. \par As has hopefully become clear, inflation is a nice idea which solves many problems of standard big bang cosmology. However, no convincing realization of inflation which does not involve unexplained small numbers has emerged (for a general discussion of this point see Ref. 120). \par It is important to distinguish between models of inflation which are self consistent and those which are not. We have shown that new inflation is not self consistent, whereas chaotic inflation is. One of the key issues involves initial conditions. In new inflation, the initial conditions required can only be obtained if the inflaton field is in thermal equilibrium above the critical temperature, which however is not possible because of the density fluctuation constraints on coupling constants. \par In chaotic inflation, it can be shown that -- provided the spatial sections are flat -- a large phase space of initial conditions (much larger than is apparent from (5.35) and (5.37)) gives chaotic inflation$^{121-123)}$, whereas the probability to relax dynamically$^{124)}$ to field configurations which give new inflation (this possibility is only available in double well potentials) is negligibly small. \section{Generation and Evolution of Fluctuations} \subsection{Preliminaries} \par In this section, the origin of the primordial density perturbations required to seed galaxies will be discussed within the context of inflationary Universe models. From Chapter 3, we recall the basic reason why in inflationary cosmology a causal generation mechanism is possible: comoving scales of cosmological interest today originate inside the Hubble radius early in the de Sitter period (see Fig. 11). Hence, it is in principle possible that density perturbations on these scales can be generated by a causal mechanism at very early times. \par First, let us demonstrate why the Hubble radius $H^{-1} (t)$ is the length scale of relevance in these considerations. Consider a scalar field theory with action $$ S (\varphi) = \int d^4 x \sqrt{-g} \, \left[ {1\over 2} \partial_\mu \varphi \partial^\mu \varphi - V (\varphi) \right] \, . \eqno\eq $$ The resulting equation of motion is $$ \ddot \varphi + 3 H \dot \varphi - a^{-2} \nabla^2 \varphi = - V^\prime (\varphi) \, . \eqno\eq $$ The second term on the left hand side is the Hubble damping term, the third represents microphysics (spatial gradients). To simplify the consideration, assume that $V^\prime (\varphi) = 0$. Then, the time evolution is influenced by microphysics and gravity. For plane wave perturbations with wave number $k$, the gravitational force is proportional to $H^2 \varphi$ whereas the microphysical force is $a^{-2} k^2 \varphi$. Thus, for $ak^{-1} < H^{-1}$, i.e., on length scales smaller than the Hubble radius, microphysics dominates, whereas for $ak^{-1} > H^{-1}$, i.e., on length scales larger than the Hubble radius, the gravitational drag dominates. \par Based on the above analysis we can formulate the main idea of the fluctuation analysis in inflationary cosmology. In linear order, all Fourier modes decouple. Hence, we fix a mode with wave number $k$. There are two very different time intervals to consider. Let $t_i (k)$ be the time when the scale crosses the Hubble radius in the de Sitter phase, and $t_f (k)$ the time when it reenters the Hubble radius after inflation (see Fig. 30). The first period lasts until $t = t_i (k)$. In this time interval microphysics dominates. We shall demonstrate that quantum fluctuations generate perturbations during this period$^{35-38, 125)}$. The second time interval is $t_i (k) < t < t_f (k)$. Now microphysics is unimportant and the evolution of perturbations is determined by gravity. At $t_i (k)$, decoherence sets in; the quantum mechanical wave functional evolves as a statistical ensemble of classical configurations after this time. \smallskip \epsfxsize=9cm \epsfbox{bfig30.eps} {\baselineskip=13pt \noindent {\bf Figure 30:} Sketch (physical distance $x_p$ versus time $t$) of the evolution of two fixed comoving scales labelled $k_1$ and $k_2$ in the inflationary Universe.} \medskip It is possible to give a heuristic derivation of the shape of the spectrum of cosmological perturbations in an inflationary Universe based on simple geometrical arguments$^{35)}$. Consider first the process of generation of fluctuations. If the generation mechanism produces inhomogeneities at all times on a fixed physical wavelength with an amplitude determined by the ``Hawking temperature" $T_H$ of de Sitter space$^{126)}$, $$ T_H = {H\over{2 \pi}} \, , \eqno\eq $$ then the evolution of perturbations on different scales between when they are formed and when they leave the Hubble radius at times $t_i (k)$ is related by time translation, and the amplitude of the fluctuations when measured at time $t_i (k)$ will be independent of $k$: $$ {\delta M\over M} (k, t_i (k)) = {\cal A}_i = \, {\rm const.} \eqno\eq $$ In fact, the amplitude ${\cal A}_i$ will be given by the thermal energy associated with the Hawking temperature (5.49) divided by the ``false vacuum" energy density $\rho_0$ responsible for inflation $$ {\cal A}_i \sim {T_H^4\over \rho_0} \, . \eqno\eq $$ In Chapter 4 it was shown (see (4.68)) that the gauge invariant measure of density fluctuations evolves trivially on scales outside of the Hubble radius: it changes by a factor which only depends on the equations of state in the inflationary and past-inflationary phases. In particular, this factor is independent of $k$. Hence, using (4.68), we arrive at the conclusion that perturbations are independent of scale when they enter the Hubble radius $$ {\delta M\over M} (k, t_f (k)) = {\cal A}_f = \, {\rm const} \, . \eqno\eq $$ Hence, we have demonstrated that a scale invariant spectrum of density perturbations is a generic feature of an inflationary Universe scenario. {}From (4.67) and (4.68) if follows that $$ {\cal A}_f \sim {\cal A}_i \, {1 + w (t_f (k))\over{1 + w (t_i (k))}} \, . \eqno\eq $$ On scales which enter the Hubble radius after $t_{eq}$, $$ 1 + w (t_f (k)) = 1 \, . \eqno\eq $$ The initial value of $w = p /\rho$ during inflation can be determined by using the expressions (3.9) for energy density and pressure of the scalar field responsible for inflation. The result is $$ 1 + w (t_i (k)) = \, {\dot \varphi^2 (t_i (k))\over{\rho (t_i (k))}} \sim {H^4\over \rho_0} \, , \eqno\eq $$ where the last proportionality follows by dimensional analysis. Thus, combining (5.50-5.54) we obtain $$ {\delta M\over M} \, (k, t_f (k)) \sim 1 \, , \eqno\eq $$ which is at least four orders of magnitude too large to conform with the constraints from CMB anisotropy measurements. The above problem is known as the ``fluctuation problem"$^{127, 90)}$ and is common to most inflationary Universe models. The only known solutions involve small numbers introduced into the particle physics sector. This defeats one of the aims of inflation which is to avoid the need for unnaturally small constants. To study this problem in more detail we must turn to a quantitative analysis of the generation and evolution of fluctuations. \subsection{Quantum Generation of Fluctuations} The question of the origin of classical density perturbations from quantum fluctuations in the de Sitter phase of an inflationary Universe is a rather subtle issue. Starting from a homogeneous quantum state ({\it e.g.}, the vacuum state in the FRW coordinate frame at time $t_i$, the beginning of inflation), a naive semiclassical anaylsis would predict the absence of fluctuations since $$ < \psi | T_{\mu\nu} (x) | \psi > = {\rm const} \, . \eqno\eq $$ However, as a simple thought experiment shows, such a naive analysis is inappropriate. Imagine a local gravitational mass detector $D$ positioned close to a large mass $M$ which is suspended from a pole (see Fig. 31). The decay of an alpha particle will sever the cord (at point $T$) by which the mass is held to the pole and the mass will drop. According to the semiclassical prescription $$ G_{\mu\nu} = 8 \pi G < \psi | T_{\mu\nu} | \psi > \, , \eqno\eq $$ the metric ({\it i.e.}, the mass measured) will slowly decrease. This is obviously not what happens. The mass detector shows a signal which corresponds to one of the classical trajectories which make up the state $| \psi >$, a trajectory corresponding to a sudden drop in the gravitational force measured. \smallskip \epsfxsize=6.5cm \epsfbox{bfig31.eps} {\baselineskip=13pt \noindent{\bf Figure 31:} Sketch of the thought experiment discussed in the text.} \medskip The origin of classical density perturbations as a consequence of quantum fluctuations in a homogeneous state $| \psi >$ can be analyzed along similar lines. The quantum to classical transition is picking out$^{128-130)}$ one of the typical classical trajectories which make up the wave function of $| \psi >$. We can implement$^{131, 132)}$ the procedure as follows: Define a classical scalar field $$ \varphi_{cl} (\underline{x} , t) = \varphi_0 (t) + \delta \varphi (\underline{x} , t) \eqno\eq $$ with vanishing spatial average of $\delta \varphi$. The induced classical energy momentum tensor $T^{cl}_{\mu\nu} (\underline{x}, t)$ which is the source for the metric is given by $$ T^{cl}_{\mu\nu} (\underline{x}, t) = T_{\mu\nu} (\varphi_{cl} (\underline{x}, t) ) \, , \eqno\eq $$ where $T_{\mu\nu} \, (\varphi_{cl} (\underline{x}, t))$ is defined as the canonical energy-momentum tensor of the classical scalar field $\varphi_{cl} (\underline{x}, t)$. Unless $\delta \varphi$ vanishes, $T^{cl}_{\mu\nu}$ is inhomogeneous. For applications to chaotic inflation, we take $| \psi >$ to be a Gaussian state with mean value $\varphi_0 (t)$ $$ < \psi | \varphi^2 (\underline{x}, t) | \psi > = \varphi_0^2 (t) \, . \eqno\eq $$ Its width is taken to be the width of the vacuum state of the free scalar field theory with mass determined by the curvature of $V(\varphi)$ at $\varphi_0$. This state is used to define the Fourier transform $\delta \tilde \varphi (k,t)$ by $$ | \delta \tilde \varphi (k) |^2 = < \psi | \, | \tilde \varphi (k) |^2 \, | \psi > \, . \eqno\eq $$ The amplitude of $\delta \tilde \varphi (k)$ is identified with the width of the ground state wave function of the harmonic osciallator $\tilde \varphi (k)$. (Recall that each Fourier mode of a free scalar field is a harmonic oscillator). Note that no divergences arise in the above construction. In principle, quantum fluctuations contribute a term to $\varphi_0 (t)$; this backreaction effect has not yet been studied. The quantum corrections to (5.60) are divergent and must be regularized and renormalized (see {\it e.g.}, Ref. 133). They are the source of the stochastic driving forces in stochastic chaotic inflation. By linearizing (5.59) about $\varphi_0 (t)$ we obtain the perturbation of the energy-momentum tensor. In particular, the energy density fluctuation $\delta \tilde \rho (k)$ is given by $$ \delta \tilde \rho (k) = \dot \varphi_0 \delta \dot {\tilde \varphi} (k) + V^\prime (\varphi_0) \delta \tilde \varphi (k) \, . \eqno\eq $$ To obtain the initial amplitude ${\cal A}_i$ of (5.49), the above is to be evaluated at the time $t_i (k)$. \par The computation of the spectrum of density perturbations produced in the de Sitter phase has been reduced to the evaluation of the expectation value (5.61). First, we must specify the state $| \psi >$. (Recall that in non-Minkowski space-times there is no uniquely defined vacuum state of a quantum field theory). We pick the FRW frame of the pre-inflationary period. In this frame, the number density of particles decreases exponentially. Hence we choose $| \psi >$ to be the ground state in this frame (see Ref. 134 for a discussion of other choices). $\psi [ \tilde \varphi (\undertext{k}), t]$, the wave functional of $|\psi >$, can be calculated explicitly. It is basically the superposition of the ground state wave functions for all oscillators $$ \psi [ \tilde \varphi (\undertext{k}), t] = N \exp \left\{ - {1\over 2} (2 \pi)^{-3} a^3 (t) \int d^3 k \omega (\undertext{k}, t) | \tilde \varphi (\undertext{k})|^2 \right\} \, . \eqno\eq $$ $N$ is a normalization constant and $\omega (\undertext{k}, t) \sim H$ at $t = t_i (k)$. Hence $$ \delta \tilde \varphi (\undertext{k}, t) = (2 \pi)^{3/2} a^{-3/2} \omega (\undertext{k}, t)^{-1/2} \sim (2 \pi)^{3/2} k^{-3/2} H \, , \, t = t_i (k)\, . \eqno\eq $$ \subsection{Evolution of Fluctuations} Given the above determination of the intitial amplitude of density perturbations at the time when they leave the Hubble radius during the de Sitter phase, and the general relativistic analysis of the evolution of fluctuations discussed in Section 4.4, it is easy to evaluate the r.m.s. inhomogeneities when they reenter the Hubble radius at time $t_f (k)$. First, we combine (5.62), (5.64), (5.49), (4.10) and (4.15) to obtain $$ \left( {\delta M\over M} \right)^2 \, (k, t_i (k)) \sim k^3 \left({V^\prime (\varphi_0) \delta \tilde \varphi (k, t_i (k))\over \rho_0} \right)^2 \sim \left({V^\prime (\varphi_0) H\over \rho_0 } \right)^2 \, , \eqno\eq $$ and thus $$ {\cal A}_i \sim \, {V^\prime (\varphi_0 (t_i (k)) H\over \rho_0} \, . \eqno\eq $$ If the background scalar field is rolling slowly, then $$ V^\prime (\varphi_0 (t_i (k))) = 3 H | \dot \varphi_0 (t_i (k)) | \, . \eqno\eq $$ Combining (5.66) and (5.67) with (5.52) and (5.54) we get $$ {\delta M\over M} (k, \, t_f (k)) = {\cal A}_f \sim \, {3 H^2 | \dot \varphi_0 (t_i (k)) |\over{\dot \varphi^2_0 (t_i (k))}} = {3 H^2\over{| \dot \varphi_0 (t_i (k))}} = {3H^2\over{| \dot \varphi_0 (t_i (k))|}} \eqno\eq $$ This result can now be evaluated for specific models of inflation to find the conditions on the particle physics parameters which give a value $$ {\cal A}_f \sim 10^{-5} \eqno\eq $$ which is required if quantum fluctuations from inflation are to provide the seeds for galaxy formation and agree with the CMB anisotropy limits. For chaotic inflation with a potential $$ V (\varphi) = {1\over 2} m^2 \varphi^2 \, , \eqno\eq $$ the dynamics of $\varphi$ was analyzed in Section 5b (see in particular (5.43) and (5.44)). We have $$ \varphi (t_i (k)) \sim m_{pl} \eqno\eq $$ and hence $$ H (t_i (k)) \sim m^{-1}_{pl} m \, \varphi (t_i (k)) \sim m \, . \eqno\eq $$ Therefore, $$ {\delta M\over M} (k, t_f (k)) \sim 3 {H^2\over{| \dot \varphi_0 (t_i (k))|}} \sim 10 {m\over m_{pl}} \eqno\eq $$ which implies that $$ m \sim 10^{13} \, {\rm GeV} \eqno\eq $$ to agree with (5.70). Similarly, for a potential of the form $$ V (\varphi) = {1\over 4} \lambda \varphi^4 \eqno\eq $$ we obtain $$ {\delta M\over M} (k, \, t_f (k)) \sim 10 \cdot \lambda^{1/2} \eqno\eq $$ which requires $$ \lambda \leq 10^{-12}. \eqno\eq $$ in order not to conflict with observations. The conditions (5.74) and (5.77) require the presence of small parameters in the particle physics model. It has been shown quite generally$^{120)}$ that small parameters are required if inflation is to solve the fluctuation problem. \subsection{Discussion} \par Let us first summarize the main results of the analysis of density fluctuations in inflationary cosmology: \item{-} Quantum vacuum fluctuations in the de Sitter phase of an inflationary Universe are the source of perturbations. \item{-} The quantum perturbations decohere on scales outside the Hubble radius and can hence be treated classically. \item{-} The classical evolution outside the Hubble radius produces a large amplification of the perturbations. In fact, unless the particle physics model contains very small coupling constants, the predicted fluctuations are in excess of those allowed by the bounds on cosmic microwave anisotropies. \item{-} Inflationary Universe models generically produce a scale invariant \hfill \break Harrison-Zel'dovich spectrum $$ {\delta M\over M} (k , t_f (k) ) = {\rm const.} \eqno\eq $$ It is not hard to construct models which give a different spectrum. All that is required is a significant change in $H$ during the period of inflation. \par I have chosen to present the analysis of fluctuations in inflationary cosmology in two separate steps in order to highlight the crucial physics issues. Having done this, it is possible to step back and construct a unified analysis in which expectation values of gauge invariant variables are propagated from $t \ll t_i (k)$ to $t_f (k)$ in a consistent way$^{135, 11)}$, and in which the final values of the expectation values of quadratic operators are used to construct $T^{cl}_{\mu \nu} (\undertext{x} , t)$. Once inside the Hubble radius, the evolution of the mass perturbations is influenced by the damping effects discussed in Section 4.3, which is turn depend on the composition of the dark matter. The dominant effects are the Meszaros effect and free streaming. On scales which enter the Hubble radius before $t_{eq}$, the perturbations can only grow logarithmically in time between $t_f(k)$ and $t_{eq}$. This implies that (up to logarithmic corrections), the mass perturbation spectrum is flat for wavelengths smaller than $\lambda_{eq}$, the comoving Hubble radius at $t_{eq}$: $$ {{\delta M} \over M} (\lambda, t) \simeq {\rm const}, \,\,\,\,\,\,\, t \leq t_{eq}, \, \lambda < \lambda_{eq}, \eqno\eq $$ whereas on larger scales $$ {{\delta M} \over M} (\lambda, t) \propto \lambda^{-2}. \eqno\eq $$ Equations (5.79) and (5.80) give the power spectrum in an $\Omega = 1$ inflationary CDM model. If the dark matter is hot, then neutrino free streaming cuts off the power spectrum at $\lambda_J^{max}$ (see (3.27)). The inflationary HDM and CDM perturbation spectra are compared in Fig. 32. \smallskip \epsfxsize=9.5cm \epsfbox{bfig32.eps} {\baselineskip=13pt \noindent{\bf Figure 32:} The linear theory power spectra for inflationary CDM (upper curve) and HDM models. The horizontal axis is mass, expressed in units of solar masses. $M_{eq}$ is the mass inside the comoving horizon at $t_{eq}$, and $M_{FS}$ is the mass inside the maximal comoving neutrino free streaming volume.} \medskip \chapter{Topological Defects and Structure Formation} \section{Classification} \par In the previous section we have seen that symmetry breaking phase transitions in unified field theories arising in particle physics ({\it e.g.,} Grand Unified Theories (GUT)$^{46)}$) do not lead, in general, to inflation. In most models, the coupling constants which arise in the effective potential for the scalar field $\varphi$ driving the phase transition are too large to generate a period of slow rolling which lasts more than one Hubble time $H^{-1} (t)$. Nevertheless, there are interesting remnants of the phase transition: topological defects. \par Consider a single component real scalar field with a typical symmetry breaking potential $$ V (\varphi) = {1\over 4} \lambda (\varphi^2 - \eta^2)^2 \eqno\eq $$ Unless $\lambda \ll 1$ there will be no inflation. The phase transition will take place on a short time scale $\tau < H^{-1}$, and will lead to correlation regions of radius $\xi < t$ inside of which $\varphi$ is approximately constant, but outside of which $\varphi$ ranges randomly over the vacuum manifold ${\cal M}$, the set of values of $\varphi$ which minimizes $V(\varphi)$ -- in our example $\varphi = \pm \eta$. The correlation regions are separated by domain walls, regions in space where $\varphi$ leaves the vacuum manifold ${\cal M}$ and where, therefore, potential energy is localized. Via the usual gravitational force, this energy density can act as a seed for structure. Topological defects are familiar from solid state and condensed matter systems. Crystal defects, for example, form when water freezes or when a metal crystallizes$^{136)}$. Point defects, line defects and planar defects are possible. Defects are also common in liquid crystals. They arise in a temperature quench from the disordered to the ordered phase$^{137)}$. Vortices in $^4$He are analogs of global cosmic strings. Vortices and other defects are also produced during a quench below the critical temperature in $^3$He$^{138)}$. Finally, vortex lines also play an important role in the theory of superconductivity$^{139)}$. The analogies between defects in particle physics and condensed matter physics are quite deep. Defects form for the same reason: the vacuum manifold is topologically nontrivial. The arguments which say that in a theory which admits defects, such defects will inevitably form, are applicable both in cosmology and in condensed matter physics. Different, however, is the defect dynamics. The motion of defects in condensed matter systems is friction-dominated, whereas the defects in cosmology obey relativistic equations, second order in time derivatives, since they come from a relativistic field theory. After these general comments we turn to a classification of topological defects. We consider theories with an $n$-component order parameter $\varphi$ and with a potential energy function (free energy density) of the form (6.1) with $$ \varphi^2 = \sum\limits^n_{i = 1} \, \varphi^2_i \, . \eqno\eq $$ There are various types of local and global topological defects (regions of trapped energy density) depending on the number $n$ of components of $\varphi$. The more rigorous mathematical definition refers to the homotopy of ${\cal M}$. The words ``local" and ``global" refer to whether the symmetry which is broken is a gauge or global symmetry. In the case of local symmetries, the topological defects have a well defined core outside of which $\varphi$ contains no energy density in spite of nonvanishing gradients $\nabla \varphi$: the gauge fields $A_\mu$ can absorb the gradient, {\it i.e.,} $D_\mu \varphi = 0$ when $\partial_\mu \varphi \neq 0$, where the covariant derivative $D_\mu$ is defined by $$ D_\mu = \partial_\mu + ie \, A_\mu \, , \eqno\eq $$ $e$ being the gauge coupling constant. Global topological defects, however, have long range density fields and forces. \par Table 1 contains a list of topological defects with their topological characteristic. A ``v" markes acceptable theories, a ``x" theories which are in conflict with observations (for $\eta \sim 10^{16}$ GeV). \smallskip \epsfxsize=12cm \epsfbox{table1.eps} {\baselineskip=13pt} \par Theories with domain walls are ruled out$^{140)}$ since a single domain wall stretching across the Universe today would overclose the Universe. Local monopoles are also ruled out$^{141)}$ since they would overclose the Universe. Local textures are ineffective at producing structures (see Section 6). We now describe examples of domain walls, cosmic strings, monopoles and textures, focussing on configurations with maximal symmetry. \par \undertext{Domain walls} arise in theories with a single real order parameter and free energy density given by (6.1). The vacuum manifold of this model consists of two points $$ {\cal M} = \{ \pm \eta \} \eqno\eq $$ and hence has nontrivial zeroth homotopy group: $$ \Pi_0 ({\cal M}) \neq 1 \eqno\eq $$ (readers not familiar with homotopy groups can simply skip all of the following statements involving $\Pi_n ({\cal M})$. They are not required for an understanding of the physics). To construct a domain wall configuration with planar symmetry (without loss of generality the $y-z$ plane can be taken to be the plane of symmetry), assume that $$ \eqalign{ \varphi (x) \simeq \eta \>\>\> & x \gg \eta^{-1} \cr \varphi (x) \simeq - \eta \>\>\> & x \ll - \eta^{-1} } \eqno\eq $$ By continuity of $\varphi$, there must be an intermediate value of $x$ with $\varphi (x) = 0$. We can take this point to be $x = 0$, {\it i.e.}, $$ \varphi (0) = 0 \, . \eqno\eq $$ The set of points with $\varphi = 0$ constitute the center of the domain wall. Physically, the wall is a thin sheet of trapped energy density. The width $w$ of the sheet is given by the balance of potential energy and tension energy. Assuming that the spatial gradients are spread out over the thickness $w$ we obtain $$ w V (0) = w \lambda \eta^4 \sim {1\over w} \, \eta^2 \eqno\eq $$ and thus $$ w \sim \lambda^{-1/2} \eta^{-1} \, . \eqno\eq $$ See Fig. 33 for a sketch of this domain wall configuration. \smallskip \epsfxsize=4.5cm \epsfbox{bfig33.eps} {\baselineskip=13pt \noindent{\bf Figure 33:} A low temperature field configuration containing a domain wall. The sketch shows a plane $P$ in space. At positions with a $+$ or $-$, the value of the scalar field is $\eta$ or $- \eta$, respectively. The solid line is the intersection with $P$ of the plane of points in space with $\varphi = 0$.} \medskip A theory with a complex order parameter $(n = 2)$ admits \undertext{cosmic strings}. In this case the vacuum manifold of the model is $$ {\cal M} = S^1 \, , \eqno\eq $$ which has nonvanishing first homotopy group: $$ \Pi_1 ({\cal M}) = Z \neq 1 \, . \eqno\eq $$ A cosmic string is a line of trapped energy density which arises whenever the field $\varphi (x)$ circles ${\cal M}$ along a closed path in space ({\it e.g.}, along a circle). In this case, continuity of $\varphi$ implies that there must be a point with $\varphi = 0$ on any sheet bounded by the closed path. The points on different sheets connect up to form a line overdensity of field energy (see Fig. 34). \smallskip \epsfxsize=6cm \epsfbox{bfig34.eps} {\baselineskip=13pt \noindent{\bf Figure 34:} Sketch of the topological argument for the existence of cosmic string configurations. Given a field configuration with nontrivial winding along a circle normal to the plane of this figure, there must be a point with $\varphi = 0$ on every disk bounded by the circle. Three disks are depicted: $D$, $D'$ and $D''$, and the respective points with $\varphi = 0$ are $z$, $z'$ and $z''$. The union of all such points $z$ forms the center $z(\lambda)$ of the string.} \medskip To construct a field configuration with a string along the $z$ axis$^{43)}$, take $\varphi (x)$ to cover ${\cal M}$ along a circle with radius $r$ about the point $(x,y) = (0,0)$: $$ \varphi (r, \vartheta ) \simeq \eta e^{i \vartheta} \, , \, r \gg \eta^{-1} \, . \eqno\eq $$ This configuration has winding number 1, {\it i.e.}, it covers ${\cal M}$ exactly once. Maintaining cylindrical symmetry, we can extend (6.12) to arbitrary $r$ $$ \varphi (r, \, \vartheta) = f (r) e^{i \vartheta} \, , \eqno\eq $$ where $f (0) = 0$ and $f (r)$ tends to $\eta$ for large $r$. The width $w$ can again be found by balancing potential and tension energy. The result is identical to the result (6.9) for domain walls. For local cosmic strings, {\it i.e.}, strings arising due to the spontaneous breaking of a gauge symmetry, the energy density decays exponentially for $r \gg \eta^{-1}$. In this case, the energy $\mu$ per unit length of a string is finite and depends only on the symmetry breaking scale $\eta$ $$ \mu \sim \eta^2 \eqno\eq $$ (independent of the coupling $\lambda$). The value of $\mu$ is the only free parameter in a cosmic string model. To see how the finiteness of the mass per unit length $\mu$ comes about, consider the simplest theory admitting local strings, the Abelian Higgs model with Lagrangean $$ {\cal L} = {1\over 2} \, D_\mu \varphi D^\mu \varphi - V (\varphi) + {1\over 4} F_{\mu\nu} F^{\mu\nu} \, , \eqno\eq $$ where $\varphi$ is a complex order parameter with potential (6.1), $D_\mu$ is the gauge covariant derivative $$ D_\mu = \partial_\mu + ie \, A_\mu \, , \eqno\eq $$ the field $A_\mu$ is a U(1) gauge potential with associated field strength $$ F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu \, , \eqno\eq $$ and $e$ is the gauge coupling constant. For an order parameter configuration (6.12), the gauge fields $A_\mu$ will take on values such that $$ D_\mu \varphi \simeq 0 \>\>\> r \gg \eta^{-1} \eqno\eq $$ even though $\partial_\mu \varphi \neq 0$. Hence, the energy density decays exponentially for $r \gg \eta^{-1}$. For strings in a global theory (no gauge potential), the spatial gradient energy $(\partial_\mu \varphi)^2$ cannot be cancelled at large $r$, and hence the mass per unit length is logarithmically divergent as a function of a large $r$ cutoff. If the order parameter of the model has three components $(n = 3)$, then \undertext{monopoles} result as topological defects. The vacuum manifold is $$ {\cal M} = S^2 \eqno\eq $$ and has topology given by $$ \Pi_2 ({\cal M}) \neq 1 \, . \eqno\eq $$ Given a sphere $S$ is space, it is possible that $\varphi$ takes on values in ${\cal M}$ everywhere on $S$, and that it covers ${\cal M}$ once. By continuity, there must be a point in space in the interior of $S$ with $\varphi = 0$. This is the center of a point-like defect, the monopole. To construct a spherically symmetric monopole with the origin as its center, consider a field configuration $\varphi$ which defines a map from physical space to field space such that all spheres $S_r$ in space of radius $r \gg \eta^{-1}$ about the origin are mapped onto ${\cal M}$ (see Fig. 35): $$ \eqalign{ \varphi: \> & S_r \longrightarrow {\cal M} \cr & (r, \vartheta,\varphi) \longrightarrow (\vartheta, \varphi) \, . }\eqno\eq $$ This configuration defines a winding number one magnitude monopole. Domain walls, cosmic strings and monopoles are examples of \undertext{topological} \undertext{defects}. A topological defect has a well-defined core, a region in space with $\varphi \notin {\cal M}$ and hence $V (\varphi) > 0$. There is an associated winding number which is quantized, {\it i.e.}, it can take on only integer values. Since the winding number can only change continuously, it must be conserved, and hence topological defects are stable. Furthermore, topological defects exist for theories with global and local symmetries. Now, let us consider a theory with a four-component order parameter ({\it i.e.,} $n = 4$), and a potential given by (6.1). In this case, the vacuum manifold is $$ {\cal M} = S^3 \eqno\eq $$ and the associated topology is given by $$ \Pi_3 ({\cal M}) \neq 1 \, . \eqno\eq $$ The corresponding defects are called ``\undertext{textures}".$^{44, 45)}$ \smallskip \epsfxsize=10cm \epsfbox{bfig35.eps} {\baselineskip=13pt \noindent {\bf Figure 35:} Construction of a monopole: left is physical space, right the vacuum manifold. The field configuration $\phi$ maps spheres in space onto ${\cal M}$. However, a core region of space near the origin is mapped onto field values not in ${\cal M}$.} \medskip Textures, however, are quite different than the previous topological defects. The texture construction will render this manifest (Fig. 36). To construct a radially symmetric texture, we give a field configuration $\varphi (x)$ which maps physical space onto ${\cal M}$. The origin 0 in space (an arbitrary point which will be the center of the texture) is mapped onto the north pole $N$ of ${\cal M}$. Spheres surrounding 0 are mapped onto spheres surrounding $N$. In particular, some sphere with radius $r_c (t)$ is mapped onto the equator sphere of ${\cal M}$. The distance $r_c (t)$ can be defined as the radius of the texture. Inside this sphere, $\varphi (x)$ covers half the vacuum manifold. Finally, the sphere at infinity is mapped onto the south pole of ${\cal M}$. The configuration $\varphi (\undertext{x})$ can be parameterized by$^{60)}$ $$ \varphi (x,y,z) = \left(\cos \chi (r), \> \sin \chi (r) {x\over r}, \> \sin \chi (r) {y\over r}, \> \sin \chi (r) {z\over r} \right) \eqno\eq $$ in terms of a function $\chi (r)$ with $\chi (0) = 0$ and $\chi (\infty) = \pi$. Note that at all points in space, $\varphi (\undertext{x})$ lies in ${\cal M}$. There is no defect core. All the energy is in spatial gradient (and possibly kinetic) terms. \par In a cosmological context, there is infinite energy available in an infinite space. Hence, it is not necessary that $\chi (r) \rightarrow \pi$ as $r \rightarrow \infty$. We can have $$ \chi (r) \rightarrow \chi_{\rm max} < \pi \>\> {\rm as} \>\> r \rightarrow \infty \, . \eqno\eq $$ \smallskip \epsfxsize=12cm \epsfbox{bfig36.eps} {\baselineskip=13pt \noindent{\bf Figure 36:} Construction of a global texture: left is physical space, right the vacuum manifold. The field configuration $\phi$ is a map from space to the vacuum manifold (see text).} \medskip In this case, only a fraction $$ n_W = {\chi_{\rm max}\over \pi} - {\sin (2 \chi_{\rm max})\over{2 \pi}} \eqno\eq $$ of the vacuum manifold is covered: the winding number $n_W$ is not quantized. This is a reflection of the fact that whereas topologically nontrivial maps from $S^3$ to $S^3$ exist, all maps from $R^3$ to $S^3$ can be deformed to the trivial map. \par Textures in $R^3$ are unstable. For the configuration described above, the instability means that $r_c (t) \rightarrow 0$ as $t$ increases: the texture collapses. When $r_c (t)$ is microscopical, there will be sufficient energy inside the core to cause $\varphi (0)$ to leave ${\cal M}$, pass through 0 and equilibrate at $\chi (0) = \pi$: the texture unwinds. \par A further difference compared to topological defects: textures are relevant only for theories with global symmetry. Since all the energy is in spatial gradients, for a local theory the gauge fields can reorient themselves such as to cancel the energy: $$ D_\mu \varphi = 0 \, . \eqno\eq $$ \par Therefore, it is reasonable to regard textures as an example of a new class of defects, \undertext{semitopological defects}. In contrast to topological defects, there is no core, and $\varphi (\undertext{x}) \in {\cal M}$ for all $\undertext{x}$. In particular, there is no potential energy. In addition, the winding number is not quantized, and hence the defects are unstable. Finally, they exist only in theories with a global internal symmetry. \section{Formation} \par The Kibble mechanism$^{12)}$ ensures that in theories which admit topological or semitopological defects, such defects will be produced during a phase transition in the very early Universe. Consider a mechanical toy model, first introduced by Mazenko, Unruh and Wald$^{93)}$ in the context of inflationary Universe models, which is useful in understanding the scalar field evolution. Consider (see Fig. 37) a lattice of points on a flat table. At each point, a pencil is pivoted. It is free to rotate and oscillate. The tips of nearest neighbor pencils are connected with springs (to mimic the spatial gradient terms in the scalar field Lagrangean). Newtonian gravity creates a potential energy $V(\varphi)$ for each pencil ($\varphi$ is the angle relative to the vertical direction). $V(\varphi)$ is minimized for $| \varphi | = \eta$ (in our toy model $\eta = \pi / 2$). Hence, the Lagrangean of this pencil model is analogous to that of a scalar field with symmetry breaking potential (6.1). \smallskip \epsfxsize=10cm \epsfbox{bfig37.eps} {\baselineskip=13pt \noindent{\bf Figure 37:} The pencil model: the potential energy of a simple pencil has the same form as that of scalar fields used for spontaneous symmetry breaking. The springs connecting nearest neighbor pencils give rise to contributions to the energy which mimic spatial gradient terms in field theory.} \medskip At high temperatures $T \gg T_c$, all pencils undergo large amplitude high frequency oscillations. However, by causality, the phases of oscillation of pencils with large separation $s$ are uncorrelated. For a system in thermal equilibrium, the length $s$ beyond which phases are random is the correlation length $\xi (t)$. However, since the system is quenched rapidly, there is a causality bound on $\xi$: $$ \xi (t) < t \, , \eqno\eq $$ where $t$ is the causal horizon. The critical temperature $T_c$ is the temperature at which the thermal energy is equal to the energy a pencil needs to jump from horizontal to vertical position. For $T < T_c$, all pencils want to lie flat on the table. However, their orientations are random beyond a distance of $\xi (t)$ determined by equating the free energy gained by symmetry breaking (a volume effect) with the gradient energy lost (a surface effect). As expected, $\xi (T)$ diverges at $T_c$. Very close to $T_c$, the thermal energy $T$ is larger than the volume energy gain $E_{corr}$ in a correlation volume. Hence, these domains are unstable to thermal fluctuations. As $T$ decreases, the thermal energy decreases more rapidly than $E_{corr}$. Below the Ginsburg temperature $T_G$, there is insufficient thermal energy to excite a correlation volume into the state $\varphi = 0$. Domains of size $$ \xi (t_G) \sim \lambda^{-1} \eta^{-1} \eqno\eq $$ freeze out$^{12, 142)}$. The boundaries between these domains become topological defects. An improved version of this argument has recently been given by Zurek$^{214)}$ (see also Ref. 215). We conclude that in a theory in which a symmetry breaking phase transitions satisfies the topological criteria for the existence of a fixed type of defect, a network of such defects will form during the phase transition and will freeze out at the Ginsburg temperature. The correlation length is initially given by (6.29), if the field $\varphi$ is in thermal equilibrium before the transition. Independent of this last assumption, the causality bound implies that $\xi (t_G) < t_G$. For times $t > t_G$ the evolution of the network of defects may be complicated (as for cosmic strings) or trivial (as for textures). In any case (see the caveats of Refs. 143 and 144), the causality bound persists at late times and states that even at late times, the mean separation and length scale of defects is bounded by $\xi (t) \leq t$. Applied to cosmic strings, the Kibble mechanism implies that at the time of the phase transition, a network of cosmic strings with typical step length $\xi (t_G)$ will form. According to numerical simulations$^{145)}$, about 80\% of the initial energy is in infinite strings (strings with curvature radius larger than the Hubble radius) and 20\% in closed loops. Note that the Kibble mechanism was discussed above in the context of a global symmetry breaking scenario. As pointed out in Ref. 146, the situation is more complicated in local theories in which gauge field can cancel spatial gradients in $\varphi$ in the energy functional, and in which spatial gradients in $\varphi$ can be gauged away. Nevertheless, as demonstrated numerically (in $2 + 1$ dimensions) in Refs. 42 and 147 and shown analytically in Ref. 148, the Kibble mechanism also applies to local symmetries. \smallskip \epsfxsize=8cm \epsfbox{bfig38.eps} {\baselineskip=13pt \noindent{\bf Figure 38:} Formation of a loop by a self intersection of an infinite string. According to the original cosmic string scenario, loops form with a radius $R$ determined by the instantaneous coherence length of the infinite string network.} \medskip The evolution of the cosmic string network for $t > t_G$ is complicated (see Section 6.4). The key processes are loop production by intersections of infinite strings (see Fig. 38) and loop shrinking by gravitational radiation. These two processes combine to create a mechanism by which the infinite string network loses energy (and length as measured in comoving coordinates). It will be shown (in Section 6.4) that as a consequence, the correlation length of the string network is always proportional to its causality limit $$ \xi (t) \sim t \, . \eqno\eq $$ Hence, the energy density $\rho_\infty (t)$ in long strings is a fixed fraction of the background energy density $\rho_c (t)$ $$ \rho_\infty (t) \sim \mu \xi (t)^{-2} \sim \mu t^{-2} \eqno\eq $$ or $$ {\rho_\infty (t)\over{\rho_c (t)}} \sim G \mu \, . \eqno\eq $$ We conclude that the cosmic string network approaches a ``scaling solution" in which the statistical properties of the network are time independent if all distances are scaled to the horizon distance. Applied to textures, the Kibble mechanism implies that on all scales $r \geq t_G$, field configurations with winding number $n_W \geq n_{cr}$ are frozen in with a probability $p (n_{cr})$ per volume $r^3$. The critical winding number $n_{cr}$ is defined as the winding number above which field configurations collapse and below which they expand. Only collapsing configurations form clumps of energy which can accrete matter. The critical winding $n_{cr}$ was determined numerically in Refs. 149 \& 150 and analytically in Ref. 151 (see also Ref. 152). It is slightly larger than 0.5. The probability $p (n_{cr})$ can be determined using combinatorial arguments$^{153)}$. For $t > t_G$, any configuration on scale $\sim t$ with winding number $n_W \ge n_{cr}$ begins to collapse (before $t$, the Hubble damping term dominates over the spatial gradient forces, and the field configuration is frozen in comoving coordinates). After unwinding, $\varphi (\undertext{x})$ is homogeneous inside the horizon. The texture model thus also leads to a scaling solution: at all times $t > t_G$ there is the same probability that a texture configuration of scale $t$ will enter the horizon, become dynamical and collapse with a typical time scale $t$. \section{Topological Defects and Cosmology} Topological defects are regions in space with trapped energy density. By Newtonian gravity, these defects can act as seeds about which the matter in the Universe clusters, and hence they play a very important role in cosmology. As indicated in Table 1, theories with domain walls or with local monopoles are ruled out, and those with only local textures do not give rise to a structure formation model. As mentioned earlier, theories with domain walls are ruled out since a single wall stretching across the present Universe would overclose it. Local monopoles are also problematic since they do not interact and come to dominate the energy density of the Universe. Local textures do not exist as coherent structures with nonvanishing gradient energy since the gauge fields can always compensate scalar field gradients. \par Let us demonstrate explicitly why stable domain walls are a cosmological disaster$^{140)}$. If domain walls form during a phase transition in the early Universe, it follows by causality (see however the caveats of Refs. 143 and 144) that even today there will be at least one wall per Hubble volume. Assuming one wall per Hubble volume, the energy density $\rho_{DW}$ of matter in domain walls is $$ \rho_{DW} (t) \sim \eta^3 t^{-1} \, , \eqno\eq $$ whereas the critical density $\rho_c$ is $$ \rho_c = H^2 \, {3\over{8 \pi G}} \sim m^2_{p\ell} \, t^{-2} \, . \eqno\eq $$ Hence, for $\eta \sim 10^{16}$ GeV the ratio of (6.33) and (6.34) is $$ {\rho_{DW}\over \rho_c} \, (t) \sim \, \left({\eta\over{m_{p\ell}}} \right)^2 \, (\eta t) \sim 10^{52} \, . \eqno\eq $$ The above argument depends in an essential way on the dimension of the defect. One cosmic string per Hubble volume leads to an energy density $\rho_{cs}$ in string $$ \rho_{cs} \sim \eta^2 \, t^{-2} \, . \eqno\eq $$ Later in this section we shall see that the scaling (6.36) holds in the cosmic string model. Hence, cosmic strings do not lead to cosmological problems. On the contrary, since for GUT models with $\eta \sim 10^{16}$ GeV $$ {\rho_{cs}\over \rho_c} \sim \, \left({\eta\over m_{p \ell}} \right)^2 \sim 10^{-6} \, , \eqno\eq $$ cosmic strings in such theories could provide the seed perturbations responsible for structure formation. Theories with local monopoles are ruled out on cosmological grounds$^{141)}$ (see again the caveats of Refs. 143 and 144) for rather different reasons. Since there are no long range forces between local monopoles, their number density in comoving coordinates does not decrease. Since their contribution to the energy density scales as $a^{-3} (t)$, they will come to dominate the mass of the Universe, provided $\eta$ is sufficiently large. Theories with global monopoles$^{154, 155)}$ are not ruled out, since there are long range forces between monopoles which lead to a ``scaling solution" with a fixed number of monopoles per Hubble volume. In the following we will describe aspects of two of the promising topological defect models of structure formation, those based on cosmic strings and on global textures. The global monopole scenario is in many aspects similar to the texture theory. \section{Cosmic String Evolution and Scaling} If the evolution of the cosmic string network were trivial in the sense that all strings would only stretch as the universe expands, there would be an immediate cosmological disaster. Consider a fixed comoving volume $V$ with a string passing through. The energy in radiation decreases as $a^{-1} (t)$ while the energy in string increases as $a (t)$. Hence trivial evolution would immediately lead to a string dominated universe, a cosmological disaster. In order to study the evolution of a cosmic string network, it is neccessary to know the effective action for a string, and to study what happens when two strings cross. \par The equations of motion of a string are determined by the Nambu action $$ S = - \mu \int d \sigma d \tau \left(- \det g^{(2)}_{ab} \right)^{1/2} \> \> a, b = 0, 1\eqno\eq $$ where $g^{(2)}_{ab}$ is the world sheet metric and $\sigma$ and $\tau$ are the world sheet coordinates. In flat space-time, $\tau$ can be taken to be coordinate time, and $\sigma$ is an affine parameter along the string. In terms of the string coordinates $X^\mu (\sigma, \tau)$ and the metric $g^{(4)}_{\mu\nu}$ of the background space-time, $$ g^{(2)}_{ab} = X^\mu_{,a} X^v_{,b} g^{(4)}_{\mu\nu} \, . \eqno\eq $$ {}From general symmetry considerations, it is possible to argue that the Nambu action is the correct action. However, I shall follow Foerster$^{156)}$ and Turok$^{157)}$ and give a direct heuristic derivation. We start from a general quantum field theory Lagrangean ${\cal L}_{QFT}$. The action is $$ S = \int d^4 y {\cal L}_{QFT} \left(\varphi (y)\right)\eqno\eq $$ We assume the existence of a linear topological defect at $X^\mu (\sigma, \tau)$. The idea now is to change variables such that $\sigma$ and $\tau$ are two of the new coordinates, and to expand $S$ to lowest order in $w/R$, where $w$ is the width of the string and $R$ its curvature radius. As the other new coordinates we take the coordinates $\rho^2$ and $\rho^3$ in the normal plane to $X^\mu (\sigma, \tau)$. Thus the transformation takes the old coordinates $y^\mu (\mu = 0, 1, 2, 3)$ to new ones $\sigma^a = (\tau, \sigma, \rho^2, \rho^3)$: $$ y^\mu (\sigma^a) = X^\mu (\sigma, \tau) + \rho^i n^\mu_i (\sigma, \tau)\eqno\eq $$ where $i = 2,3$ and $n^\mu_i$ are the basis vectors in the normal plane to the string world sheet. The measure transforms as $$ \int d^4 y = \int d \sigma d \tau d \rho^2 d \rho^3 (\det M_a^\mu)\eqno\eq $$ with $$ M^\mu_a = \, {\partial y^\mu\over{\partial \sigma^a}} = \, \pmatrix{\partial X^\mu/\partial (\sigma, \tau)\cr n^\mu_i\cr} + O (\rho)\, .\eqno\eq $$ The determinant can easily be evaluated using the following trick $$ \det M^\mu_a = \, \left( - \det \eta_{\mu \nu} M^\mu_a M^\nu_b \right)^{1/2} \equiv \sqrt{- \det D_{ab}}\eqno\eq $$ $$ D = \, \pmatrix{{\partial x^\mu\over{\partial (\sigma, \tau)}} \, {\partial X^\nu\over{\partial (\sigma, \tau)}} \eta_{\mu \nu} & {\partial X^\mu\over {\partial (\sigma , \tau)}} n^\nu_b \eta_{\mu\nu}\cr {\partial X^\mu\over{\partial (\sigma, \tau)}} n^\nu_a \eta_{\mu\nu} & n^\mu_a n^\nu_b \eta_{\mu\nu}\cr} = \pmatrix{X^\mu_{,a} X^\nu_{,b} \eta_{\mu\nu} & 0\cr 0 & \delta_{ab}\cr} + 0 \, \left({w\over R}\right) \eqno\eq $$ Hence $$ \eqalign{S &= \int d \sigma d \tau \left( - \det g^{(2)}_{ab} \right)^{1/2} \int d \rho^2 d \rho^3 {\cal L} (y (\sigma, \tau, \rho^2, \rho^3)) + O \left({w\over R} \right)\cr &= - \mu \int d \sigma d \tau \left( - \det g^{(2)}_{ab} \right)^{1/2} + O \left({w\over R}\right)\, . }\eqno\eq $$ $- \mu$ is the integral of {\cal L} in the normal plane of $X$. To first order in $w/R$, it equals the integral of $-{\cal H}$; hence it is the mass per unit length. \par This derivation of the Nambu action is instructive as it indicates a method for calculating corrections to the equations of motion of the string when extra fields are present, \eg\ for superconducting cosmic strings. It also gives a way of calculating the finite thickness corrections to the equations of motion which will be important at cusps (see below). \par In flat space-time we can consistently choose $\tau = t, \dot x \cdot x^\prime = 0$ and $\dot x^2 + x^{\prime^2} = 0$. The equations of motion derived from the Nambu action then become $$ \ddot {\undertext{x}} - \undertext{x}^{\prime\prime} = 0\, . \eqno\eq $$ where $\prime$ indicates the derivative with respect to $\sigma$. The general solution can be decomposed into a left moving and a right moving mode$^{158)}$ $$ \undertext{x} (t, \sigma) = {1\over 2} \, \left[ \undertext{a} (\sigma - t) + \undertext{b} (\sigma + t ) \right] \eqno\eq $$ The gauge conditions imply $$ \dot {\undertext{a}}^2 = \dot {\undertext{b}}^2 = 1\eqno\eq $$ For a loop, $\undertext{x} (\sigma, t)$ is periodic and hence the time average of $\dot {\undertext{a}}$ and $\dot {\undertext{b}}$ vanish. $\dot {\undertext{a}}$ and $\dot {\undertext{b}}$ are hence closed curves on the unit sphere with vanishing average. Two such curves generically intersect if they are continuous. An intersection corresponds to a point with $\undertext{x}^\prime = 0$ and $\dot {\undertext{x}} = 1$. Such a point moving at the speed of light is called a cusp. $\dot {\undertext{x}} (\sigma, t)$ need not be continuous. Points of discontinuity are called kinks. Note that both cusps and kinks will be smoothed out by finite thickness effects$^{159)}$. \par The Nambu action does not describe what happens when two strings hit. This process has been studied numerically for both global$^{160)}$ and local$^{161)}$ strings. The authors of these papers set up scalar field configurations corresponding to two strings approaching one another and evolve the complete classical scalar field equations. The result of the analysis is that strings do not cross but exchange ends, provided the relative velocity is smaller than 0.9. Thus, by self intersecting, an infinite string will split off a loop (Fig. 38). An important open problem is to understand this process analytically. For a special value of the coupling constant Ruback$^{162)}$ has given a mathematical explanation (see also Shellard and Ruback in Ref. 161). \par There are two parts to the nontrivial evolution of the cosmic string network. Firstly, loops are produced by self intersections of infinite strings. Loops oscillate due to the tension and slowly decay by emitting gravitational radiation. Combining the two steps we have a process by which energy is transferred from the cosmic string network to radiation. \par There are analytical indications that a stable ``scaling solution" (already described in Section ) for the cosmic string network exists. In the scaling solution, on the order of 1 infinite string segment crosses every Hubble volume. The correlation length $\xi (t)$ of an infinite string is thus of the order $t$. A heuristic argument for the scaling solution is due to Vilenkin$^{5)}$. Take $\tilde \nu (t)$ to be the mean number of infinite string segments per Hubble volume. Then the energy density in infinite strings is $$ \rho_\infty (t) = \mu \tilde \nu (t) t^{-2} \eqno\eq $$ The number of loops $n(t)$ produced per unit volume is proportional to the square of $\tilde \nu$, since it takes two string segments to generate a string intersection. Hence, $$ {d n (t)\over{dt}} = c \tilde \nu^2 t^{-4} \eqno\eq $$ where $c$ is a constant of the order $1$. Conservation of energy in strings gives $$ {d \rho_\infty (t)\over{dt}} + {3\over{2 t}} \, \rho_\infty (t) = - c^\prime \mu t \, {dn\over{dt}} = - c^\prime \mu \tilde \nu^2 t^{-3} \eqno\eq $$ or, written as an equation for $\tilde \nu (t)$ $$ \tilde {\dot \nu} - \, {\tilde \nu\over{2 t}} = - cc^\prime \tilde \nu^2 t^{- 1}\eqno\eq $$ Thus if $\tilde \nu \gg 1$ then $\tilde {\dot \nu} < 0$ while if $\tilde \nu \ll 1$ then $\tilde {\dot \nu} > 0$. Hence there will be a stable solution with $\tilde \nu \sim 1$. The precise value of $\tilde \nu$ must be determined in numerical simulations. These simulations are rather difficult because of the large dynamic range required and due to singularities which arise in the evolution equations near cusps. In the radiation dominated epoch, $\tilde \nu$ is still uncertain by a factor of about 10. The first results were reported in Ref. 163. More recent results are due three groups. Bennett and Bouchet$^{164)}$ and Allen and Shellard$^{165)}$ are converging on a value $10 < \tilde \nu < 20$, whereas Albrecht and Turok$^{166)}$ obtain a value which is about 100. \par The scaling solution for the infinite strings implies that the network of strings looks the same at all times when scaled to the Hubble radius. This should also imply that the distribution of cosmic string loops is scale invariant in the same sense. At present, however, there is no convincing evidence from numerical simulations that this is really the case. \par A scaling solution for loops implies that the distribution of $R_i (t)$, the radius of loops at the time of formation, is time independent after dividing by $t$. To simplify the discussion, I shall assume that the distribution is monochromatic, \ie\ $$ R_i (t)/t = \alpha\, . \eqno\eq $$ Based on Fig. 38, we expect $\alpha \sim 1$. The numerical simulations$^{164-166)}$, however, now give $\alpha < 10^{-2}$. This is due to the fact that there is a lot of small scale structure on the long strings, and that the typical scale of loop production is not determined by the overall curvature radius of the long strings, but rather by the typical lengths of the small scale structure. \par {}From the scaling solution (6.50) for the infinite strings we can derive the scaling solution for loops. We assume that the energy density in long strings -- inasmuch as it is not redshifted -- must go into loops. $\beta$ shall be a measure for the mean length $\ell$ in a loop of ``radius" $R$ $$ \ell = \beta R\, . \eqno\eq $$ If per expansion time and Hubble volume about 1 loop of radius $R_i (t)$ is produced, then we know that the number density in physical coordinates of loops of radius $R_i (t)$ is $$ n (R_i (t), t) = ct^{-4}\eqno\eq $$ with a constant $c$ which can be calculated from (6.50), (6.54) and (6.55). Neglecting gravitational radiation, this number density simply redshifts $$ n (R,t) = \, \left({z (t)\over{z (t_f (R))}} \right)^3 n (R, t_f (R))\, , \eqno\eq $$ where $t_f (R)$ is the time when loops of radius $R$ are formed. Isolating the $R$ dependence, we obtain $$ n (R, t) \sim R^{-4} z (R)^{-3}\eqno\eq $$ where $z (R)$ is the redshift at time $t=R$. We have the following special cases: $$ \eqalign{n (R, t) \sim R^{-5/2} t^{-3/2} \qquad & t < t_{eq}\cr n (R, t) \sim R^{-5/2} t_{eq}^{1/2} t^{-2} \qquad & t > t_{eq} \, , \, t_f (R) < t_{eq}\cr n (R, t) \sim R^{-2} t^{-2} \qquad & t > t_{eq} \, , \, t_f (R) > t_{eq} \, .}\eqno\eq $$ \par The proportionality constant $c$ is $$ c = {1\over 2} \beta^{-1} \alpha^{-2} \tilde \nu\eqno\eq $$ (see \eg\ Ref. 167). In deriving (6.60) it is important to note that $n (R_i (t), t) dR_i$ is the number density of loops in the radius interval $[R_i , R_i + dR_i]$. Hence, in the radiation dominated epoch $$ n (R, t) = \nu R^{- 5/2} t^{-3/2} \eqno\eq $$ with $$ \nu = {1\over 2} \beta^{-1} \alpha^{1/2} \tilde \nu \, .\eqno\eq $$ \par {}From (6.62) we can read off the uncertainties in $\nu$ based on the uncertainties in the numerical results. Both $\alpha^{1/2}$ and $\tilde \nu$ are determined only up to one order of magnitude. Hence, any quantitative results which depend on the exact value of $\nu$ are rather uncertain. \par Gravitational radiation leads to a lower cutoff in $n (R, t)$. Loops with radius smaller than this cutoff were all formed at essentially the same time and hence have the same number density. Thus, $n (R)$ becomes flat. The power in gravitational radiation $P_G$ can be estimated using the quadrupole formula$^{168)}$. For a loop of radius $R$ and mass $M$ $$ P_G = {1\over 5} G < \dot{\ddot Q} \dot{\ddot Q} > \, , \eqno\eq $$ where $Q$ is the quadrupole moment, $Q \sim MR^2$, and since the frequency of oscillation is $\omega = R^{-1}$ $$ P_G \sim G (MR^2)^2 \omega^6 \sim (G \mu) \mu \, . \eqno\eq $$ \par Even though the quadrupole approximation breaks down since the loops move relativistically, (6.64) gives a good order of magnitude of the power of gravitational radiation. Improved calculations give$^{169)}$ $$ P_G = \gamma (G \mu) \mu\eqno\eq $$ with $\gamma \sim 50$. (6.55) and (6.65) imply that $$ \dot R = \tilde \gamma G \mu\eqno\eq $$ with $\tilde \gamma \equiv \gamma/\beta \sim 5$ (using $\beta \simeq 10$). Note that the rate of decrease is constant. Hence, $$ R (t) = R_i - (t - t_i) \tilde \gamma G \mu\eqno\eq $$ and the cutoff loop radius is $$ R_c \sim \tilde \gamma G \mu t_i\, . \eqno\eq $$ \par Let us briefly summarize the scaling solution \item{1)} At all times the network of infinite strings looks the same when scaled by the Hubble radius. A small number of infinite string segments cross each Hubble volume and $\rho_\infty (t)$ is given by (6.50). \item{2)} There is a distribution of loops of all sizes $0 \le R < t$. Assuming scaling for loops, then $$ n (R, t) = \nu R^{-4} \, \left({z (t)\over{z (R)}}\right)^3 \, , \> R \> \epsilon \> [ \tilde \gamma G \mu t, \alpha t]\eqno\eq $$ where $\alpha^{-1} R$ is the time of formation of a loop of radius $R$. Also $$ n (R, t) = n ( \tilde \gamma G \mu t, t) \, , \> R < \tilde \gamma G \mu t\, . \eqno\eq $$ Although the qualitative characteristics of the cosmic string scaling solution are well established, the quantitative details are not. The main reason for this is the fact that the Nambu action breaks down at kinks and cusps. However, kinks and cusps inevitably form and are responsible for the small scale structure on strings. In fact, coarse graining by integrating out the small scale structure may give an equation of state for strings which deviates from that of a Nambu string$^{170)}$. Attempts at understanding the small scale structure on strings are at present under way$^{171)}$. \section{Cosmic Strings and Structure Formation} The starting point of the structure formation scenario in the cosmic string theory is the scaling solution for the cosmic string network, according to which at all times $t$ (in particular at $t_{eq}$, the time when perturbations can start to grow) there will be a few long strings crossing each Hubble volume, plus a distribution of loops of radius $R \ll t$ (see Fig. 39). \smallskip \epsfxsize=6.5cm \epsfbox{bfig39.eps} {\baselineskip=13pt \noindent {\bf Figure 39.} Sketch of the scaling solution for the cosmic string network. The box corresponds to one Hubble volume at arbitrary time $t$.} \medskip The cosmic string model admits three mechanisms for structure formation: loops, filaments, and wakes. Cosmic string loops have the same time averaged field as a point source with mass$^{172)}$ $$ M (R) = \beta R \mu \, , \eqno\eq $$ $R$ being the loop radius and $\beta \sim 2 \pi$. Hence, loops will be seeds for spherical accretion of dust and radiation. For loops with $R \leq t_{eq}$, growth of perturbations in a model dominated by cold dark matter starts at $t_{eq}$. Hence, the mass at the present time will be $$ M (R, \, t_0) = z (t_{eq}) \beta \, R \mu \, . \eqno\eq $$ In the original cosmic string model$^{47, 48, 57)}$ it was assumed that loops dominate over wakes. In this case, the theory could be normalized ({\it i.e.}, $\mu$ could be determined) by demanding that loops with the mean separation of clusters $d_{cl}$ (from the discussion in Section 6.4 it follows that the loop radius $R (d_{cl})$ is determined by the mean separation) accrete the correct mass, {\it i.e.}, that $$ M (R (d_{cl}), t_0) = 10^{14} M_{\odot} \, . \eqno\eq $$ This condition yields$^{57)}$ $$ \mu \simeq 10^{32} {\rm GeV}^2 \eqno\eq $$ Thus, if cosmic strings are to be relevant for structure formation, they must arise due to a symmetry breaking at energy scale $\eta \simeq 10^{16}$GeV. This scale happens to be the scale of unification (GUT) of weak, strong and electromagnetic interactions. It is tantalizing to speculate that cosmology is telling us that there indeed was new physics at the GUT scale. \smallskip \epsfxsize=12cm \epsfbox{bfig40.eps} {\baselineskip=13pt \noindent{\bf Figure 40.} Sketch of the mechanism by which a long straight cosmic string $S$ moving with velocity $v$ in transverse direction through a plasma induces a velocity perturbation $\Delta v$ towards the wake. Shown on the left is the deficit angle, in the center is a sketch of the string moving in the plasma, and on the right is the sketch of how the plasma moves in the frame in which the string is at rest.} \medskip The second mechanism involves long strings moving with relativistic speed in their normal plane which give rise to velocity perturbations in their wake$^{173)}$. The mechanism is illustrated in Fig. 40: space normal to the string is a cone with deficit angle$^{174)}$ $$ \alpha = 8 \pi G \mu \, . \eqno\eq $$ If the string is moving with normal velocity $v$ through a bath of dark matter, a velocity perturbation $$ \delta v = 4 \pi G \mu v \gamma \eqno\eq $$ [with $\gamma = (1 - v^2)^{-1/2}$] towards the plane behind the string results. At times after $t_{eq}$, this induces planar overdensities, the most prominent ({\it i.e.}, thickest at the present time) and numerous of which were created at $t_{eq}$, the time of equal matter and radiation$^{58, 59, 63)}$. The corresponding planar dimensions are (in comoving coordinates) $$ t_{eq} z (t_{eq}) \times t_{eq} z (t_{eq}) v \sim (40 \times 40 v) \, {\rm Mpc}^2 \, . \eqno\eq $$ The thickness $d$ of these wakes can be calculated using the Zel'dovich approximation$^{63)}$. The result is $$ d \simeq G \mu v \gamma (v) z (t_{eq})^2 \, t_{eq} \simeq 4 v \, {\rm Mpc} \, . \eqno\eq $$ \par Wakes arise if there is little small scale structure on the string. In this case, the string tension equals the mass density, the string moves at relativistic speeds, and there is no local gravitational attraction towards the string. In contrast, if there is small scale structure on strings, then the string tension $T$ is smaller$^{170)}$ than the mass per unit length $\mu$ and the metric of a string in $z$ direction becomes$^{175)}$ $$ ds^2 = (1 + h_{00}) (dt^2 - dz^2 - dr^2 - (1 - 8G \mu) r^2 dy^2 ) \eqno\eq $$ with $$ h_{00} = 4G (\mu - T) \ln \, {r\over r_0} \, , \eqno\eq $$ $r_0$ being the string width. Since $h_{00}$ does not vanish, there is a gravitational force towards the string which gives rise to cylindrical accretion, thus producing filaments. As is evident from the last term in the metric (6.79), space perpendicular to the string remains conical, with deficit angle given by (6.75). However, since the string is no longer relativistic, the transverse velocities $v$ of the string network are expected to be smaller, and hence the induced wakes will be shorter and thinner. Which of the mechanisms -- filaments or wakes -- dominates is determined by the competition between the velocity induced by $h_{00}$ and the velocity perturbation of the wake. The total velocity is$^{175)}$ $$ u = - {2 \pi G (\mu - T)\over{v \gamma (v)}} - 4 \pi G \mu v \gamma (v) \, , \eqno\eq $$ the first term giving filaments, the second producing wakes. Hence, for small $v$ the former will dominate, for large $v$ the latter. By the same argument as for wakes, the most numerous and prominent filaments will have the distinguished scale $$ t_{eq} z (t_{eq}) \times d_f \times d_f \eqno\eq $$ where $d_f$ can be calculated using the Zel'dovich approximation$^{216)}$. The cosmic string model predicts a scale-invariant spectrum of density perturbations, exactly like inflationary Universe models but for a rather different reason. Consider the {\it r.m.s.} mass fluctuations on a length scale $2 \pi k^{-1}$ at the time $t_H (k)$ when this scale enters the Hubble radius. From the cosmic string scaling solution it follows that a fixed ({\it i.e.}, $t_H (k)$ independent) number $\tilde v$ of strings of length of the order $t_H (k)$ contribute to the mass excess $\delta M (k, \, t_H (k))$. Thus $$ {\delta M\over M} \, (k, \, t_H (k)) \sim \, {\tilde v \mu t_H (k)\over{G^{-1} t^{-2}_H (k) t^3_H (k)}} \sim \tilde v \, G \mu \, . \eqno\eq $$ Note that the above argument predicting a scale invariant spectrum will hold for all topological defect models which have a scaling solution, in particular also for global monopoles and textures. The amplitude of the {\it r.m.s.} mass fluctuations (equivalently: of the power spectrum) can be used to normalize $G \mu$. Since today on galaxy cluster scales $$ {\delta M\over M} (k, \, t_0) \sim 1 \, , \eqno\eq $$ the growth rate of fluctuations linear in $a(t)$ yields $$ {\delta M\over M} \, (k, \, t_{eq}) \sim 10^{-4} \, , \eqno\eq $$ and therefore, using $\tilde v \sim 10$, $$ G \mu \sim 10^{-5} \, . \eqno\eq $$ A big advantage of the cosmic string model over inflationary Universe models is that HDM is a viable dark matter candidate. Cosmic string loops survive free streaming, as discussed in Section 3.4, and can generate nonlinear structures on galactic scales, as discussed in detail in Refs. 61 and 62. Accretion of hot dark matter by a string wake was studied in Ref. 63. In this case, nonlinear perturbations develop only late. At some time $t_{nl}$, all scales up to a distance $q_{\rm max}$ from the wake center go nonlinear. Here $$ q_{\rm max} \sim G \mu v \gamma (v) z (t_{eq})^2 t_{eq} \sim 4 v \, {\rm Mpc} \, , \eqno\eq $$ and it is the comoving thickness of the wake at $t_{nl}$. Demanding that $t_{nl}$ corresponds to a redshift greater than 1 leads to the constraint $$ G \mu > 5 \cdot 10^{-7} \, . \eqno\eq $$ Note that in a cosmic string and hot dark matter model, wakes form nonlinear structures only very recently. Accretion onto loops and small scale structure on the long strings provide two mechanisms which may lead to high redshift objects such as quasars and high redshift galaxies. The first mechanism has recently been studied in Ref. 217. The power spectra in the cosmic string models with CDM and HDM are obviously different on scales smaller than the maximal neutrino free streaming length (3.27). Recent calculations$^{176, 177)}$ of the power spectra are shown in Fig. 41. \smallskip \epsfxsize=12cm \epsfbox{bfig41.eps} {\baselineskip=13pt \noindent{\bf Figure 41:} Power spectra for cosmic string HDM and CDM theories (dashed curves), compared to those for inflationary HDM and CDM models (solid curves). In each case, the top curve is for CDM, the bottom one for HDM. Note that there is substantial power on small scales in the cosmic string HDM theory.} \medskip \section{Global Textures and Structure Formation} The starting point of the texture scenario of structure formation$^{60)}$ is the scaling solution for textures: at any time $t$, there is a fixed probability $p (n_w) \, dn_w$ that the scalar field configuration over a Hubble volume covers between $n_w$ and $n_w + dn_w$ of the vacuum manifold, {\it i.e.}, we have a texture with winding number in the interval $[n_w, \, n_w + dn_w ]$ entering the Hubble radius. The dynamics of a texture is easy to understand. Consider the spherically symmetric texture configuration of (6.24) with $\chi (r)$ increasing from 0 to $\chi_{\rm max}$ over a distance $d$. If $d$ is larger than the Hubble radius, then the Hubble damping term dominates the equation of motion for $\varphi$ and the field configuration is frozen in. Once the Hubble radius $t$ catches up with $d$, the microphysical forces become dominant and the texture field begins to evolve. \smallskip \epsfxsize=9cm \epsfbox{bfig42.eps} {\baselineskip=13pt \noindent{\bf Figure 42:} A sketch of the forces acting on a spherically symmetric texture configuration and which cause unwinding in case (a) in which the winding number is larger than the critical winding, and dissipation if the winding is smaller than its critical value (case (b)). $r$ is the distance from the center of the texture, and the vertical axis shows the value of the $\chi$ field.} \medskip The evolution of $\varphi$ tends to minimize the field energy. Consider first large distances from the texture center. The spatial gradient energy can be decreased by having $\chi_{\rm max}$ increase (if $\chi_{\rm max} > \chi_c$) or decrease (if $\chi_{\rm max} < \chi_c$) (see Figure 42). The winding associated with $\chi_c$ is called the critical winding $n_c$ (see (6.26)). For a single texture in an infinite volume we would expect $$ \chi_c = {\pi\over 2} \, (i.e., \, n_c = 0.5) \, . \eqno\eq $$ For realistic textures there will be a ``finite volume cutoff" determined by the separation of textures. A semi-analytical analysis and numerical simulations give$^{149-151)}$ $$ n_c \simeq 0.6 \, . \eqno\eq $$ If $n_W < n_c$, then the field configuration will relax to a trivial one. No localized energy concentrations will be generated, and we cannot speak of a ``texture." However, if $n_W > n_c$ the field evolution will be more interesting. At large $r$, $\chi (r)$ will increase. In addition, the radius $r (\chi)$ where $\chi$ takes on a fixed value $\chi$ tends to decrease, since this leads to a concentration of gradient energies over a smaller region. Hence, the field configuration will contract (see Fig. 42), with increasing total winding number. Eventually, close to $r = 0$ there is sufficient tension energy for $\varphi$ to be able to leave the vacuum manifold and jump from $\chi = 0$ to $\chi = \pi$. This is the texture unwinding event. After unwinding, energy is radiated radially in the form of Goldstone bosons. In the texture model it is the contraction of the field configuration which leads to density perturbations$^{178)}$. At the time when the texture enters the horizon, an isocurvature perturbation is established: the energy density in the scalar field is compensated by a deficit in radiation. However, the contraction of the scalar field configuration leads to a clumping of gradient and kinetic energy at the center of the texture (Fig. 43). This, in turn, provides the seed perturbations which cause dark matter and radiation to collapse in a spherical manner$^{179, 180)}$. \smallskip \epsfxsize=9cm \epsfbox{bfig43.eps} {\baselineskip=13pt \noindent {\bf Figure 43}: A sketch of the density perturbation produced by a collapsing texture. The left graph shows the time evolution of the field $\chi (r)$ as a function of radius $r$ and time (see (5.18)). The contraction of $\chi (r)$ leads to a spatial gradient energy perturbation at the center of the texture, as illustrated on the right. The energy is denoted by $\rho$. Solid lines denote the initial time, dashed lines are at time $t + \Delta t$, and dotted lines correspond to time $t + 2 \Delta t$, where $\Delta t$ is a fraction of the Hubble expansion time (the typical time scale for the dynamics).} \medskip As in the cosmic string model, also in the global texture scenario the length scale of the dominant structures is the comoving Hubble radius at $t_{eq}$. Textures generated at $t_{eq}$ are the most numerous, and the perturbations induced by them have the most time to grow. As mentioned in the previous subsection, the texture model predicts a scale-invariant spectrum of density perturbations. Hence, in order to differentiate topological defect models from inflationary scenarios, and to distinguish between different topological defect theories, we need statistics which are not determined by the power spectrum alone. We need statistics which are sensitive to the non-random phases of topological defect models. One such statistic is the genus curve$^{181)}$. For a surface $S$ embedded in $R^3$, the genus $g$ is $$ g (S) = {\rm \# \, of \, holes \, of} \, S - \, {\rm \# \, of \, disconnected \, components \, of} \, S + 1 \, . \eqno\eq $$ The genus $g$ can now be evaluated for the isodensity surface $S (\rho)$, the surface of points in space with density equal to $\rho$. The curve $$ g (\rho) = g (S (\rho)) \eqno\eq $$ is the genus curve. To reduce numerical errors, $g$ can also be evaluated based on a cell decomposition of the volume. Now, $g (n)$ is the genus of the boundary of the cell complex in which each cell contains more than $n$ galaxies. In this case, the genus is simply $$ g = 1 - {1\over 2} (V-E-F) \eqno\eq $$ where $V,E,F$ are the number of vertices, edges, and faces of the polygonal surface, respectively. \smallskip \epsfxsize=11.5cm \epsfbox{bfig44.eps} {\baselineskip=13pt \noindent{\bf Figure 44:} Comparison of the genus curve (genus as a function of galaxy density) of different toy models of structure formation. Except for the Gaussian model, all theories have the same linear power spectrum. The `filament', `wake' and `texture' toy models are based on laying down at random linear, planar and spherical overdense regions of galaxies. Thus, the figure demonstrates that the genus statistic is able to distinguish between theories with different topologies but identical power spectra. The `CDM' model predictions are computed from linear theory, and the `Poisson' model is obtained by randomly distributing galaxies. See the senior thesis by Aguirre$^{182)}$ for further details.} \medskip As shown in Fig. 44, the genus statistic is able to distinguish between models with the same power spectrum but different phase correlations and topology$^{182)}$. For a texture toy model, the genus curve is mostly negative, for a cosmic string wake model, it is predominately positive. The differences compared to a random phase inflationary model are statistically significant. The differences shown in Fig. 44 will only be apparent in large-scale samples of galaxies, {\it i.e.}, on scales exceeding the comoving radius at $t_{eq}$. Such samples should, however, become available in the near future, and at that point genus curve and other statistics sensitive to non-random phases should become a powerful tool for distinguishing the predictions of the different models of structure formation. A final word concerning textures: since they are short-lived, only CDM is a viable dark matter candidate in the context of this structure formation scenario. \chapter{Cosmic Microwave Background Anisotropies} As mentioned in Section 3, the near-isotropy of the CMB is the strongest evidence in support of the cosmological principle. By the same reasoning, any density inhomogeneities in the early Universe will give rise to CMB anisotropies. Since our present theories of galaxy formation are based on the gravitational instability scenario, they predict such inhomogeneities. The CMB temperature fluctuations probe the structure of space at $t_{rec}$, the time of last scattering, a time when the density perturbations still have a small amplitude and can be analyzed in linear theory. Hence, a study of CMB anisotropies will yield a lot of constraints for structure formation models. The information gained will be robust, {\it i.e.}, independent of the uncertainties of nonlinear gravitational and hydrodynamical effects, but it will deal only with large scales (comparable or larger than the comoving horizon at $t_{rec}$). In this section we shall give a brief overview of the theory of CMB anisotropies and summarize some recent observational results. \section{Basics} As illustrated in Fig. 45, there are three main sources of CMB anisotropies. The first are gravitational potential perturbations at $t_{rec}$ which lead to fluctuations of the surface of last scattering. This produces deviations in the light travel time between last scattering and detection, and -- given that the photons have the same temperature on the surface of last scattering -- to temperature fluctuations for the observer. {\baselineskip=13pt \noindent{\bf Figure 45:} Space-time plot sketching the origin of CMB temperature anisotropies. The surface labelled $T_{rec}$ is the last scattering surface. ${\cal O}$ is the observer at the present time measuring photons $\gamma$ impinging from directions in the sky separated by an angle $\theta$. The shaded area labelled $C$ is the world volume of a local overdensity, leading to distortions of geodesics. Possible velocities of observer and emitter are indicated by ${\vec v}_o$ and ${\vec v}_e$, respectively.} \medskip The second source is due to gravitational perturbations along the line of sight which lead to deviations of the geodesics and hence to temperature differences. A Newtonian way of understanding this effect is to consider a photon passing through a large mass concentration. On the way towards the center, the photon is falling into a potential well and acquires a blueshift, whereas on its way out it is redshifted. In an expanding background, this redshift does not exactly cancel the initial blueshift, and a temperature fluctuation results. The third source contributing to CMB anisotropies are peculiar velocities on the surface of last scattering and of the observer. The peculiar motion of the earth gives rise to a dipole anisotropy$^{183)}$ $$ {\delta T\over T} \big|_{\rm dipole} \simeq \, 10^{-3} \eqno\eq $$ Peculiar velocities induce temperature fluctuations by means of the Doppler effect. For linear adiabatic density perturbations in a matter dominated Universe, the line of sight contributions to $\delta T/T$ can be written as total time derivative and thus reduces to a contribution from the surface of last scattering and can be simply combined with the potential fluctuations at $t_{rec}$. This is the case first studied by Sachs and Wolfe, and the combined effect is now called the Sachs-Wolfe effect$^{184)}$. An analysis of the Sachs-Wolfe effect reveals a very simple relationship between temperature fluctuations $\delta T/T \, (\vartheta)$ on an angular scale $\vartheta$ and the magnitude of density perturbations on the corresponding lengths scale $\lambda (\vartheta)$, where at last scattering $\lambda (\vartheta)$ equals the distance subtended by two light rays with angular separation $\vartheta$ (see Fig. 45). A simple derivation$^{11, 185)}$ of this relationship makes use of the gauge invariant theory of cosmological perturbations described in Section 4.4. The starting point is the phase space distribution function $f (x^\alpha, \, p_i)$ which would be a function of $p / T$ exclusively in the absence of inhomogeneities. In the presence of inhomogeneities, the deviation of $f$ from homogeneity is associated with temperature fluctuations: $$ f (x^\alpha , \, p_i) = \bar f (p / \bar T + \delta T) \, , \eqno\eq $$ where $\bar T$ is the average temperature and $\bar f (p / T)$ is the background phase space density. The phase space distribution function satisfies the collisionless Boltzmann equation $$ {d x^\alpha\over{d \eta}} \, {\partial f\over{\partial x^\alpha}} + {d p_i\over{d \eta}} \, {\partial f\over{\partial p_i}} = 0 \eqno\eq $$ where, as in Section 4.4, the variable $\eta$ denotes conformal time. This equation can be integrated along the perturbed geodesics which are given by $$ {d p_\alpha\over{d \eta}} = 2p {\partial \Phi\over{\partial x^\alpha}} \eqno\eq $$ and $$ {d x^i\over{d \eta}} = l^i (1 + 2 \Phi), \eqno\eq $$ with $$ l^i = - {1\over p} \, p_i \eqno\eq $$ and $$ p^2 = p_i p_i \, . \eqno\eq $$ Inserting these relations into the Boltzman equation gives $$ \left( {\partial\over{\partial \eta}} + l^i \partial_i \right) \, {\delta T\over T} = - 2 l^i \partial_i \Phi \, . \eqno\eq $$ Since in the matter dominated period $\partial_\eta \Phi = 0$ we can rewrite (7.8) as $$ \left( {\partial\over{\partial_\eta}} + l^i_i \right) \, \left({\delta T\over T} + 2 \Phi \right) = 0 \, , \eqno\eq $$ which implies that $$ {\delta T\over T} + 2 \Phi = \, {\rm const} \eqno\eq $$ along the perturbed geodesics. For isothermal primordial perturbations $( {\delta T\over T} \, (t_{rec}) = 0)$, the result (7.10) implies that $$ {\delta T\over T} (\eta_0) = 2 \Phi (\eta_{rec}) + l^i v_i (\eta_{rec}) \eqno\eq $$ whereas for primordial adiabatic perturbations (vanishing initial entropy perturbations) $$ {\delta T\over T} (\eta_0) = {1\over 3} \Phi (\eta_{rec}) + l^i v_i (\eta_{rec}) \, , \eqno\eq $$ {\it i.e.}, the combination of initial curvature fluctuations and line of sight effects leads to a partial cancellation of the anisotropy. The second term on the {\rm r.h.s.} of (7.11) and (7.12) is the Doppler term, and it arises from a determination of the constant in (7.10) based on considering the initial conditions at $t_{rec}$. Since $\Phi$ is constant both between $t_{eq}$ and $t_{rec}$ and while outside the Hubble radius, and since $$ \Phi (t_H) \sim {\delta \rho\over \rho} \, (t_H) \eqno\eq $$ at Hubble radius crossing $t_H$, our results imply that (modulo Doppler terms) for adiabatic perturbations $$ {\delta T\over T} (\vartheta, \, t_0) = {1\over 3} \Phi (\lambda (\vartheta), \, t_{eq}) \sim {1\over 3} \, {\delta M\over M} \, (\lambda (\vartheta), \, t_H)) \, . \eqno\eq $$ We conclude that the spectrum of primordial mass perturbations can be normalized by CMB anisotropy detections. For a scale invariant spectrum of density perturbations, the {\it r.m.s.} temperature fluctuations are predicted to be independent of $\vartheta$ on angular scales larger than the Hubble radius at $t_{rec}$ (between 1 and 2 degrees). \section{Specific Signatures} All theories of structure formation give rise to Sachs-Wolfe type temperature fluctuations given by (7.13) and (7.14). In topological defect models there are, in addition, specific signatures which cannot be described in a linear perturbative analysis. As described in Section 6.5, space perpendicular to a long straight cosmic string is conical with deficit angle given by (6.75). Consider now CMB radiation approaching an observer in a direction normal to the plane spanned by the string and its velocity vector (see Fig. 46). Photons arriving at the observer having passed on different sides of the string will obtain a relative Doppler shift which translates into a temperature discontinuity of amplitude$^{186)}$ $$ {\delta T\over T} = 4 \pi G \mu v \gamma (v) \, , \eqno\eq $$ where $v$ is the velocity of the string. Thus, the distinctive signature for cosmic strings in the microwave sky are line discontinuities in $T$ of the above magnitude. {\baselineskip=13pt \noindent{\bf Figure 46:} Sketch of the Kaiser-Stebbins effect by which cosmic strings produce linear discontinuities in the CMB. Photons $\gamma$ passing on different sides of a moving string $S$ (velocity $v$) towards the observer ${\cal O}$ receive a relative Doppler shift due to the conical nature of space perpendicular to the string (deficit angle $\alpha$).} \medskip Given ideal maps of the CMB sky it would be easy to detect strings. However, real experiments have finite beam width. Taking into account averaging over a scale corresponding to the beam width will smear out the discontinuity, and it turns out to be surprisingly hard to distinguish the predictions of the cosmic string model from that of inflation-based theories using quantitative statistics which are easy to evaluate analytically, such as the kurtosis of the spatial gradient map of the CMB$^{187)}$. Textures produce a distribution of hot and cold spots on the CMB sky with typical size of several degrees$^{188)}$. This signature is much easier to see in CMB maps. The mechanism which produces these hot and cold spots in the CMB is illustrated in Fig. 47. {\baselineskip=13pt \noindent{\bf Figure 47:} Space-time diagram of a collapsing texture. The unwinding occurs at the point $TX$. The shaded areas correspond to overdense regions. Photons like $\gamma_1$ are redshifted, those like $\gamma_2$ are blueshifted.} \medskip Photons arriving at the observer having passed through a texture as in the case of the ray $\gamma_1$ in Fig. 47 will be redshifted relative to the average photons since they have to climb out of a potential well, whereas those in orientation $\gamma_2$ will be blueshifted since they fall into a potential well. Taking into account reionization produced by texture collapse gives an amplitude of $\delta T/T$ of$^{189, 190)}$ $$ {\delta T\over T} \sim 0.06 \times 16 \pi G \eta^2 \, . \eqno\eq $$ A number of about ten hot and cold spots of angular scale 10$^\circ$ is predicted by the texture model. Theories of structure formation can now be normalized from CMB anisotropy data and from large-scale structure considerations. An inflationary model with CDM yields agreement between these two normalizations provided$^{191)}$ $$ b \simeq 1 \, , \eqno\eq $$ where $b$ is the bias factor determining the ratio of fractional mass to light perturbations on a scale of 8h$^{-1}$ Mpc. $$ {\delta L\over L} \Big|_{8 {\rm h}^{-1}{\rm Mpc}} = b \, {\delta M\over M} \Big|_{8 {\rm h}^{-1} {\rm Mpc}} \, . \eqno\eq $$ However, agreement between galaxy and cluster correlation properties seem to require$^{192)}$ $$ b \sim 2 \, . \eqno\eq $$ Normalizations of the texture model from large-scale structure and CMB observations$^{189, 190, 193)}$ require a bias $$ b \sim 3 \, , \eqno\eq $$ whereas for cosmic strings the two normalizations agree well. Based both on numerical simulations and analytical calculations, a normalization of the cosmic string model from the COBE CMB anisotropy data gives$^{194, 195)}$ $$ G \mu = (1.3 \pm 0.5) 10^{-6} \, . \eqno\eq $$ \section{Experimental Results} Over the past couple of years there has been a spectacular breakthrough on the observational front. The DMR experiment on the COBE satellite$^{196)}$ has produced a temperature map of the entire sky with beam width of 7$^\circ$, which shows a clear detection of CMB anisotropies. Independent confirmation has come from two 5$^\circ$ experiments, FIRAS$^{197)}$ which has mapped 1/4 of the sky, and the Tenerife experiment$^{198)}$ which surveyed a strip of 70$^\circ$ length in right ascension at a declination 40$^\circ$. The FIRAS data cross correlate very well with the COBE results, and there is even good agreement in the location of a pronounced feature in the Tenerife map with a that of a comparable feature in the two-year COBE maps. In addition, there are many small angular scale experiments which have detected anisotropies. A partial list of observational results is given in Table 2 . In this table, ``Angular Scale" denotes the beam width, the ``results for $\delta T/T$" stands for the variance of $\delta T$ computed from the CMB maps, ``cover" indicates the area of the sky mapped. MAX 1 and MAX 2 denote two separate MAX measurements of $\delta T/T$, one in a region of the sky $\mu \, {\rm Peg}$, the second near GUM. OVRO 1 is the first Owens Valley experiment, a measurement near the North Galactic Cap, the second is a ring survey. The large numer of anisotropy experiments which have announced detections of temperature fluctuations since April 1992 indicates the rapid progress in this field. To a first approximation, the present experimental results are in agreement with the predictions of a scale invariant spectrum of density perturbations. A popular way to show the results is to expand $T(\undertext{n})$ in spherical harmonics $$ T (\undertext{n}) = \sum_l \sum_{m = -l}^l a_{lm} Y_{lm} (\undertext{n}) ,\eqno\eq $$ where $\undertext{n}$ is a unit vector on the sky, and to calculate the temperature correlation function $$ < T (\undertext{n}_1) T (\undertext{n}_2 ) > = {1\over{4 \pi}} \sum\limits_l (2 l + 1) C_l P_l (\undertext{n}_1 \cdot \undertext{n}_2 ) \eqno\eq $$ where $$ < a^\ast_{lm} \, a_{l^\prime m^\prime} > = C_l \delta_{ll^\prime} \delta_{mm^\prime} \, . \eqno\eq $$ For a power spectrum of density perturbations $$ P (k) \sim k^n \eqno\eq $$ the prediction for the Sachs-Wolfe contribution to $\delta T$ is $$ l^2 C_l \sim l^{n-1} \eqno\eq $$ on scales larger than the Hubble radius at $t_{rec}$ ({\it i.e.}, for small values of $l$). A direct comparison between theory and experiment is complicated by two effects: the Doppler contribution to $\delta T/T$ creates a peak in the $l^2 C_l$ curve at values of $l$ which correspond to wavelengths comparable to the Hubble radius at $t_{rec}$, whose amplitude depends strongly on the ionization history of the Universe. Reionization also leads to a decrease in $C_l$ for large $l$. The COBE results combined with Tenerife observations favor$^{198)}$ a value of $n$ larger than what is predicted by simple inflationary models. However, the error bars are large and the difference is not (yet) statistically significant. At present there is the intriguing puzzle as to why the signal of certain small scale experiments is larger than the upper limit of other observations at the same angular scale elsewhere in the sky. A search for possible non-Gaussian features in the CMB sky will have high priority in the next years. \midinsert \bigskip \settabs 5\columns \centerline{{\bf TABLE 2}:~ CMB Anisotropy Results} \medskip \+ Experiment & Angular Scale & Result for ${\delta T\over T}$ & Cover & Location \cr \vskip 10pt \+ & & & & \cr \vskip 12pt \+ COBE-DMR$^{196)}$ & 7$^\circ$ & 1.1 $\pm$ 0.2 & 4$\pi$ & space \cr \vskip 12pt \+ Tenerife$^{198)}$ & 5.6$^\circ$ & 1.7 $\pm$ 0.4 & 350 deg$^2$ & ground \cr \vskip 12pt \+ FIRS$^{197)}$ & 4$^\circ$ & 1 - 3 & $\pi$ & balloon \cr \vskip 12pt \+ SK93$^{199)}$ & 1.45$^\circ$ & 1.4 $\pm$ 0.5 & & ground \cr \vskip 12pt \+ SP91$^{200)}$ & 1.4$^\circ$ & 1.1 $\pm$ 0.5 & 13.8 deg$^2$ & ground \cr \vskip 12pt \+ ARGO$^{201)}$ & 1$^\circ$ & 2.2 $\pm$ 0.8 & & balloon \cr \vskip 12pt \+ Python$^{202)}$ & 0.75$^\circ$ & 3 & 8 deg$^2$ & ground \cr \vskip 12pt \+ MAX 1$^{203)}$ & 0.5$^\circ$ & $< 3$ & & balloon \cr \vskip 12pt \+ MAX 2$^{204)}$ & 0.5$^\circ$ & 4.9 $\pm$ 0.8 & 6 deg$^2$ & balloon\cr \vskip 12pt \+ MSAM$^{205)}$ & 0.47$^\circ$ & 1.6 $\pm$ 0.4 & 6 deg$^2$ & balloon\cr \vskip 12pt \+ White Dish$^{206)}$ & 0.2$^\circ$ & $< 2.3$ & & ground \cr \vskip 12pt \+ OVRO 1$^{207)}$ & 1.8$^\prime$ & $<1.9$ & 0.03 deg$^2$ & ground \cr \vskip 12pt \+ OVRO 2$^{208)}$ & 1.8$^\prime$ & 3.4 $\pm$ 1.1 & 0.1 deg$^2$ & ground \cr \bigskip \endinsert \bigskip \bigskip \chapter{Modern Cosmology and Planck Scale Physics} \medskip \section{Introduction} Through its implications for very early Universe cosmology, Planck scale physics (and specifically string theory) might well have directly observable consequences for the physical world. The aim of this chapter is to explore some possibilities of how this may occur. As was explained in Chapter 5, standard particle physics models do not yield a convincing realization of inflation since in this context, inflation requires a fundamental scalar field with a reasonably flat potential (in order to have inflation) and with very small coupling constants (in order that quantum fluctuations present during inflation do not lead to CMB temperature anisotropies in excess of those recently detected. Such potentials are not generic in particle physics models. The first challenge from cosmologists to Planck scale physics is therefore to provide a generic mechanism for inflation. It may be that Planck scale physics predicts the type of scalar field potentials for which successful inflation results. Another possibility is that Planck scale physics leads to a realization of inflation which does not involve scalar fields. A possible scenario for this is suggested in Section 2. Finally, it may be that Planck scale physics leads to a solution of the homogeneity and flatness problems which does not require inflation. Standard and modern cosmology are plagued by an internal inconsistency. They predict that the Universe started at a ``Big Bang" singularity with infinite curvature and matter temperature. However, it is known that the physics on which the standard cosmological model is built must break down at very high temperature and curvature. Therefore, the second challenge for Planck scale physics is to find a solution to the singularity problem. Two very different scenarios in which this may happen are suggested in Sections 2 and 3. Finally, Planck scale physics (string theory as a concrete example) allows us to ask questions about the physical world which cannot be posed in standard physics. For example, is there a dynamical mechanism which singles out a Universe in which three space and one time dimensions are observable? One mechanism in the context of string theory will be reviewed in Section 3. I will review two very different approaches to Planck scale cosmology. The first is an attempt to incorporate Planck scale effects on the space-time structure by writing down an effective action for the space-time metric. It will be shown that a class of effective actions exists whose solutions have a less singular structure. More specifically, all homogeneous and isotropic solutions are nonsingular (see Section 2). In Section 3, I will summarize some aspects of string cosmology and indicate how in the context of string theory the cosmological singularities can be avoided. A dynamical mechanism which explains why at most three-spatial dimensions are large (and thus observable) is suggested. \section{A Nonsingular Universe} \subsection{Motivation} Planck scale physics will generate corrections to the Einstein action which determines the dynamics of the space-time metric $g_{\mu\nu}$. This can be seen by considering the effective action obtained by integrating out quantum matter fields in the presence of a dynamical metric, by calculating first order perturbative quantum gravity effects, or by studying the low energy effective action of a Planck scale unified theory such as string theory. The question we wish to address in this section is whether it is possible to construct a class of effective actions for gravity which have improved singularity properties and which predict inflation, with the constraint that they give the correct low curvature limit. What follows is a summary of recent work$^{41, 218, 219)}$ in which we have constructed an effective action for gravity in which all solutions with sufficient symmetry are nonsingular. The theory is a higher derivative modification of the Einstein action, and is obtained by a constructive procedure well motivated in analogy with the analysis of point particle motion in special relativity. The resulting theory is asymptotically free in a sense which will be specified below. A possible objection to our approach is that near a singularity quantum effects will be important and therefore a classical analysis is doomed to fail. This argument is correct in the usual picture in which at high curvatures there are large fluctuations and space-time becomes more like a ``quantum foam." However, in our theory, at high curvature space-time becomes highly regular and thus a classical analysis of space-time is self-consistent. The property of asymptotic freedom is essential in order to reach this conclusion. Our aim is to construct a theory with the property that the metric $g_{\mu\nu}$ approaches the de Sitter metric $g_{\mu\nu}^{DS}$, a metric with maximal symmetry which admits a geodesically complete and nonsingular extension, as the curvature $R$ approaches the Planck value $R_{pl}$. Here, $R$ stands for any curvature invariant. Naturally, from our classical considerations, $R_{pl}$ is a free parameter. However, if our theory is connected with Planck scale physics, we expect $R_{pl}$ to be set by the Planck scale. {\baselineskip=13pt \noindent{\bf Figure 48:} Penrose diagrams for collapsing Universe (left) and black hole (right) in Einstein's theory (top) and in the nonsingular Universe (bottom). C, E, DS and H stand for contracting phase, expanding phase, de Sitter phase and horizon, respectively, and wavy lines indicate singularities.} \medskip If successful, the above construction will have some very appealing consequences. Consider, for example, a collapsing spatially homogeneous Universe. According to Einstein's theory, this Universe will collapse in finite proper time to a final ``big crunch" singularity (top left Penrose diagram of Figure 48). In our theory, however, the Universe will approach a de Sitter model as the curvature increases. If the Universe is closed, there will be a de Sitter bounce followed by re-expansion (bottom left Penrose diagram in Figure 48). Similarly, in our theory spherically symmetric vacuum solutions would be nonsingular, i.e., black holes would have no singularities in their centers. The structure of a large black hole would be unchanged compared to what is predicted by Einstein's theory (top right, Figure 48) outside and even slightly inside the horizon, since all curvature invariants are small in those regions. However, for $r \rightarrow 0$ (where $r$ is the radial Schwarzschild coordinate), the solution changes and approaches a de Sitter solution (bottom right, Figure 48). This would have interesting consequences for the black hole information loss problem. To motivate our effective action construction, we turn to a well known analogy, point particle motion in the theory of special relativity. \subsection{An Analogy} The transition from the Newtonian theory of point particle motion to the special relativistic theory transforms a theory with no bound on the velocity into one in which there is a limiting velocity, the speed of light $c$ (in the following we use units in which $\hbar = c = 1$). This transition can be obtained$^{41)}$ by starting with the action of a point particle with world line $x(t)$: $$ S_{\rm old} = \int dt {1\over 2} \dot x^2 \, , \eqno\eq $$ and adding$^{220)}$ a Lagrange multiplier which couples to $\dot x^2$, the quantity to be made finite, and which has a potential $V(\varphi)$: $$ S_{\rm new} = \int dt \left[ {1\over 2} \dot x^2 + \varphi \dot x^2 - V (\varphi) \right] \, .\eqno\eq $$ {}From the constraint equation $$ \dot x^2 = {\partial V\over{\partial \varphi}} \, , \eqno\eq $$ it follows that $\dot x^2$ is limited provided $V(\varphi)$ increases no faster than linearly in $\varphi$ for large $|\varphi|$. The small $\varphi$ asymptotics of $V(\varphi)$ is determined by demanding that at low velocities the correct Newtonian limit results: $$ \eqalign{V (\varphi) \sim \varphi^2 \> & {\rm as} \> |\varphi| \rightarrow 0 \, , \cr V (\varphi) \sim \varphi \> & {\rm as} \> |\varphi| \rightarrow \infty \, . } \eqno\eq $$ Choosing the simple interpolating potential $$ V (\varphi) = {2 \varphi^2\over{1 + 2 \varphi}} \, , \eqno\eq $$ the Lagrange multiplier can be integrated out, resulting in the well-known action $$ S_{\rm new} = {1\over 2} \int dt \sqrt{1 - \dot x^2} \eqno\eq $$ for point particle motion in special relativity. \subsection{Construction} Our procedure for obtaining a nonsingular Universe theory$^{41)}$ is based on generalizing the above Lagrange multiplier construction to gravity. Starting from the Einstein action, we can introduce a Lagrange multiplier $\varphi_1$ coupled to the Ricci scalar $R$ to obtain a theory with limited $R$: $$ S = \int d^4 x \sqrt{-g} (R + \varphi_1 \, R + V_1 (\varphi_1) ) \, , \eqno\eq $$ where the potential $V_1 (\varphi_1)$ satisfies the asymptotic conditions (8.4). However, this action is insufficient to obtain a nonsingular gravity theory. For example, singular solutions of the Einstein equations with $R=0$ are not effected at all. The minimal requirements for a nonsingular theory is that \underbar{all} curvature invariants remain bounded and the space-time manifold is geodesically complete. Implementing the limiting curvature hypothesis$^{221)}$, these conditions can be reduced to more manageable ones. First, we choose one curvature invariant $I_1 (g_{\mu\nu})$ and demand that it be explicitely bounded, i.e., $|I_1| < I_1^{pl}$, where $I_1^{pl}$ is the Planck scale value of $I_1$. In a second step, we demand that as $I_1 (g_{\mu\nu})$ approaches $I_1^{pl}$, the metric $g_{\mu\nu}$ approach the de Sitter metric $g^{DS}_{\mu\nu}$, a definite nonsingular metric with maximal symmetry. In this case, all curvature invariants are automatically bounded (they approach their de Sitter values), and the space-time can be extended to be geodesically complete. Our approach is to implement the second step of the above procedure by another Lagrange multiplier construction$^{41)}$. We look for a curvature invariant $I_2 (g_{\mu\nu})$ with the property that $$ I_2 (g_{\mu\nu}) = 0 \>\> \Leftrightarrow \>\> g_{\mu\nu} = g^{DS}_{\mu\nu} \, , \eqno\eq $$ introduce a second Lagrange multiplier field $\varphi_2$ which couples to $I_2$ and choose a potential $V_2 (\varphi_2)$ which forces $I_2$ to zero at large $|\varphi_2|$: $$ S = \int d^4 x \sqrt{-g} [ R + \varphi_1 I_1 + V_1 (\varphi_1) + \varphi_2 I_2 + V_2 (\varphi_2) ] \, , \eqno\eq $$ with asymptotic conditions (8.4) for $V_1 (\varphi_1)$ and conditions $$ \eqalign{V_2 (\varphi_2) & \sim {\rm const} \>\> {\rm as} \> | \varphi_2 | \rightarrow \infty \cr V_2 (\varphi_2) & \sim \varphi^2_2 \>\> {\rm as} \> |\varphi_2 | \rightarrow 0 \, ,} \eqno\eq $$ for $V_2 (\varphi_2)$. The first constraint forces $I_2$ to zero, the second is required in order to obtain the correct low curvature limit. These general conditions are reasonable, but not sufficient in order to obtain a nonsingular theory. It must still be shown that all solutions are well behaved, i.e., that they asymptotically reach the regions $|\varphi_2| \rightarrow \infty$ of phase space (or that they can be controlled in some other way). This must be done for a specific realization of the above general construction. \subsection{Specific Model} At the moment we are only able to find an invariant $I_2$ which singles out de Sitter space by demanding $I_2 = 0$ provided we assume that the metric has special symmetries. The choice $$ I_2 = (4 R_{\mu\nu} R^{\mu\nu} - R^2 + C^2)^{1/2} \, , \eqno\eq $$ singles out the de Sitter metric among all homogeneous and isotropic metrics (in which case adding $C^2$, the Weyl tensor square, is superfluous), all homogeneous and anisotropic metrics, and all radially symmetric metrics. We choose the action$^{41)}$ $$ S = \int d^4 x \sqrt{-g} \left[ R + \varphi_1 R - (\varphi_2 + {3\over{\sqrt{2}}} \varphi_1) I_2^{1/2} + V_1 (\varphi_1) + V_2 (\varphi_2) \right] \eqno\eq $$ with $$ V_1 (\varphi_1) = 12 \, H^2_0 {\varphi^2_1\over{1 + \varphi_1}} \left( 1 - {\ln (1 + \varphi_1)\over{1 + \varphi_1}} \right) \eqno\eq $$ $$ V_2 (\varphi_2) = - 2 \sqrt{3} \, H^2_0 \, {\varphi^2_2\over{1 + \varphi^2_2}} \, . \eqno\eq $$ The general equations of motion resulting from this action are quite messy. However, when restricted to homogeneous and isotropic metrics of the form $$ ds^2 = dt^2 - a (t)^2 (dx^2 + dy^2 + dz^2) \, , \eqno\eq $$ the equations are fairly simple. With $H = \dot a / a$, the two $\varphi_1$ and $\varphi_2$ constraint equations are $$ H^2 = {1\over{12}} V^\prime_1 \eqno\eq $$ $$ \dot H = - {1\over{2\sqrt{3} }} V^\prime_2 \, , \eqno\eq $$ and the dynamical $g_{00}$ equation becomes $$ 3 (1 - 2 \varphi_1) H^2 + {1\over 2} (V_1 + V_2) = \sqrt{3} H (\dot \varphi_2 + 3 H \varphi_2) \, . \eqno\eq $$ The phase space of all vacuum configurations is the half plane $\{ (\varphi_1 \geq 0, \, \varphi_2) \}$. Equations (8.16) and (8.17) can be used to express $H$ and $\dot H$ in terms of $\varphi_1$ and $\varphi_2$. The remaining dynamical equation (8.18) can then be recast as $$ {d \varphi_2\over{d \varphi_1}} = - {V_1^{\prime\prime}\over{4 V^\prime_2}} \, \left[ - \sqrt{3} \varphi_2 + (1 - 2\varphi_1) - {2\over{V^\prime_1}} (V_1 + V_2) \right] \, . \eqno\eq $$ The solutions can be studied analytically in the asymptotic regions and numerically throughout the entire phase space. The resulting phase diagram of vacuum solutions is sketched in Fig. 49 (for numerical results, see the second article in Ref. 41). The point $(\varphi_1, \, \varphi_2) = (0,0)$ corresponds to Minkowski space-time $M^4$, the regions $|\varphi_2 | \rightarrow \infty$ to de Sitter space. As shown, all solutions either are periodic about $M^4$ or else they asymptotically approach de Sitter space. Hence, all solutions are nonsingular. This conclusion remains unchanged if we add spatial curvature to the model. {\baselineskip=13pt \noindent{\bf Figure 49:} Phase diagram of the homogeneous and isotropic solutions of the nonsingular Universe. The asymptotic regions are labelled by A, B, C and D, flow lines are indicated by arrows.} \medskip One of the most interesting properties of our theory is asymptotic freedom$^{41)}$, i.e., the coupling between matter and gravity goes to zero at high curvatures. It is easy to add matter (e.g., dust or radiation) to our model by taking the combined action $$ S = S_g + S_m \, , \eqno\eq $$ where $S_g$ is the gravity action previously discussed, and $S_m$ is the usual matter action in an external background space-time metric. We find$^{41))}$ that in the asymptotic de Sitter regions, the trajectories of the solutions in the $(\varphi_1, \, \varphi_2)$ plane are unchanged by adding matter. This applies, for example, in a phase of de Sitter contraction when the matter energy density is increasing exponentially but does not affect the metric. The physical reason for asymptotic freedom is obvious: in the asymptotic regions of phase space, the space-time curvature approaches its maximal value and thus cannot be changed even by adding an arbitrary high matter energy density. Naturally, the phase space trajectories near $(\varphi_1, \, \varphi_2) = (0,0)$ are strongly effected by adding matter. In particular, $M^4$ ceases to be a stable fixed point of the evolution equations. \subsection{Connection with Dilaton Gravity} The low energy effective actions for the space-time metric in 4 dimensions which come from string theory are only known perturbatively. They contain higher derivative terms, but not if the exact same form as the ones used in our construction. The connection between our limiting curvature construction and string theory-motivated effective actions is more apparent in two space-time dimensions$^{218, 219)}$. The most general renormalizable Lagrangian for string-induced dilaton gravity is $$ {\cal L} = \sqrt{-g} [ D(\varphi) R + G (\varphi) (\nabla \varphi)^2 + H (\varphi) ] \, , \eqno\eq $$ where $\varphi (x,t)$ is the dilaton. In two space-time dimensions, the kinetic term for $\varphi$ can be eliminated, resulting in a Lagrangian (in terms of rescaled fields) of the form $$ {\cal L} = \sqrt{-g} [ D(\varphi) R + V (\varphi) ] \, . \eqno\eq $$ We can now apply the limiting curvature construction to find classes of potentials for which the theory has nonsingular black hole$^{218)}$ and cosmological$^{219)}$ solutions. In the following, we discuss the nonsingular two-dimensional black hole. To simplify the algebra, the dilaton is redefined such that $$ D (\varphi) = {1\over \varphi} \, . \eqno\eq $$ The most general static metric can be written as $$ ds^2 = f (r) dt^2 - g (r) dr^2 \eqno\eq $$ and the gauge choice $$ g (r) = f (r)^{-1} \eqno\eq $$ is always possible. The variational equations are $$ f^\prime = - V (\varphi) {\varphi^2\over \varphi^\prime} \, , \eqno\eq $$ $$ \left( {\varphi^\prime\over \varphi^2} \right)^\prime = 0 \eqno\eq $$ and $$ \varphi^{-2} R = {\partial V\over{\partial \varphi}} \, , \eqno\eq $$ where a prime denotes the derivative with respect to $r$. Equation (8.27) can be integrated to find (after rescaling $r$) $$ \varphi = {1\over{Ar}} \, . \eqno\eq $$ To give the correct large $r$ behavior for the metric, we need to impose that $$ f (r) \rightarrow 1 - {2m\over r} \>\>\> {\rm as} \> r \rightarrow \infty \, . \eqno\eq $$ {}From (8.26) this leads to the asymptotic condition $$ V (\varphi) \rightarrow 2 m A^3 \varphi^2 \>\>\> {\rm as} \> \varphi \rightarrow 0 \, . \eqno\eq $$ The limiting curvature hypothesis requires that $R$ be bounded as $\varphi \rightarrow \infty$. From (8.28) this implies $$ V (\varphi) \rightarrow {2\over{\ell^2 \varphi}} \>\>\> {\rm as} \> \varphi \rightarrow \infty \, , \eqno\eq $$ where $\ell$ is a constant which determines the limiting curvature. As an interpolating potential we can choose $$ V (\varphi) = {2 m A^3 \varphi^2\over{1+ m A^3 \ell^2 \varphi^3}} \, , \eqno\eq $$ which allows (8.26) to be integrated explicitly$^{218)}$ to obtain $f(r)$. The resulting metric coefficient $f(r)$ describes a nonsingular black hole with a single horizon at $r \simeq 2m$. The metric is indistinguishable from the usual Schwarzschild metric until far inside of the horizon, where our $f(r)$ remains regular and obtains vanishing derivative at $r = 0$, which allows for a geodesically complete extension of the manifold. \subsection{Discussion} We have shown that a class of higher derivative extensions of the Einstein theory exist for which many interesting solutions are nonsingular. This class of models is very special. Most higher derivative theories of gravity have, in fact, much worse singularity properties than the Einstein theory. What is special about this class of theories is that they are obtained using a well motivated Lagrange multiplier construction which implements the limiting curvature hypothesis. We have shown that \item{\rm i)} all homogeneous and isotropic solutions are nonsingular$^{41)}$ \item{\rm ii)} the two-dimensional black holes are nonsingular$^{218)}$ \item{\rm iii)} nonsingular two-dimensional cosmologies exist$^{219)}$. \noindent We also have evidence that four-dimensional black holes and anisotropic homogeneous cosmologies are nonsingular$^{222)}$. By construction, all solutions are de Sitter at high curvature. Thus, the theories automatically have a period of inflation (driven by the gravity sector in analogy to Starobinsky inflation$^{39)}$) in the early Universe. A very important property of our theories is asymptotic freedom. This means that the coupling between matter and gravity goes to zero at high curvature, and might lead to an automatic suppression mechanism for scalar fluctuations. In two space-time dimensions, there is a close connection between dilaton gravity and our construction. In four dimensions, the connection between fundamental physics and our class of effective actions remains to be explored. In particular, it would be nice to investigate the connection between our limiting curvature construction and the `pre-big-bang cosmology' scenario proposed on the basis of dilaton gravity in Ref. 119. \section{Aspects of String Cosmology} \subsection{Motivation} In the previous section we studied effective actions for the space-time metric which might arise in the intermediate energy regime of a fundamental theory such as string theory. However, it is also of interest to explore the predictions of string theory which depend specifically on the ``stringy" aspects of the theory and which are lost in any field theory limit. It is to a description of a few of the string-specific cosmological aspects to which we turn in this section. \subsection{Implications of Target Space Duality} Target space duality$^{223)}$ is a symmetry specific to string theory. As a simple example, consider a superstring background in which all spatial dimensions are toroidally compactified with equal radii. Let $R$ denote the radius of the torus. The spectrum of string states is spanned by oscillatory modes which have energies independent of $R$, by momentum modes whose energies $E_n$ (with integer $n$) are $$ E_n = {n\over R} \, , \eqno\eq $$ and by winding modes with energies $E^\prime_m$ ($m$ integer) $$ E^\prime_m = mR \, . \eqno\eq $$ Target space duality is a symmetry between two superstring theories, one on a background with radius $R$, the other on a background of radius $1/R$, under which winding and momentum modes are interchanged. Target space duality has interesting consequences for string cosmology$^{224)}$. Consider a background with adiabatically changing $R(t)$. While $R(t) \gg 1$, most of the energy in thermal equilibrium resides in the momentum modes. The position eigenstates $|x >$ are defined as in quantum field theory in terms of the Fourier transform of the momentum eigenstates $|p >$ $$ |x > = \sum\limits_p e^{i x \cdot p} |p > \, . \eqno\eq $$ However, for $R (t) \ll 1$, most of the energy flows into winding modes, and it takes much less energy to measure the ``dual distance" $| \tilde x >$ than $|x >$, where $$ | \tilde x > = \sum\limits_w e^{i \tilde x \cdot w} | w > \eqno\eq $$ is defined in terms of the winding modes $| w>$. We conclude that target space duality in string theory leads to a minimum physical length in string cosmology. As $R(t)$ decreases below 1, the measured length starts to increase again. This could lead to a bouncing or oscillating cosmology$^{224)}$. It is well known that for strings in thermal equilibrium there is a maximal temperature, the Hagedorn temperature$^{225)}$. Target space duality implies that in thermal equilibrium the temperature in an adiabatically varying string background begins to decrease once $R(t)$ falls below 1: $$ T \left({1\over R} \right) = T(R) \, . \eqno\eq $$ Thus, the $T(R)$ curve in string cosmology is nonsingular and very different from its behavior in standard cosmology. For further discussions of the thermodynamics of strings see, e.g., Refs. 226 and 227 and references therein. \subsection{Strings and Space-Time Dimensionality} Computations$^{224)}$ using the microcanonical ensemble show that for all spatial directions compactified at large total energy $E$, the entropy $S$ is proportional to $E$: $$ S = \beta_H E \, , \eqno\eq $$ with $\beta_H$ denoting the inverse of the Hagedorn temperature $T_H$. Thus, the $E(R)$ curve in string cosmology is very different from the corresponding curve in standard cosmology. For large $R \gg 1$, most of the energy in a gas of strings in thermal equilibrium will flow into momentum modes, and the thermodynamics will approach that of an ideal gas of radiation for which $$ E (R) \sim {1\over R} \, . \eqno\eq $$ By duality, for small $R$ $$ E (R) \sim R \, . \eqno\eq $$ If, however, for some reason the string gas falls out of equilibrium, the $E(R)$ curve will look very different. Starting at $R= 1$ with a temperature approximately equal to $T_H$, a large fraction of the energy will reside in winding modes. If these winding modes cannot annihilate, thermal equilibrium will be lost, and the energy in winding modes will increase linearly in $R$, and thus for large $R$: $$ E (R) \sim R \, . \eqno\eq $$ Newtonian intuition tells us that out of equilibrium winding modes with an energy relation (8.42) will prevent the background space from expanding$^{224}$. The equation of state corresponding to a gas of straight strings is $$ p = - {1\over N} \rho \eqno\eq $$ where $p$ and $\rho$ denote pressure and energy density, respectively, and $N$ is the number of spatial dimensions. According to standard general relativity, an equation of state with negative pressure will lead to more rapid expansion of the background. It turns out that the Newtonian intuition is the correct one and that general relativity gives the wrong answer$^{228)}$. At high densities, the specific stringy effects -- in particular target space duality -- become crucial. The Einstein action violates duality. In order to restore duality, it is necessary to include the dilaton in the effective action for the string background. The action for dilaton gravity is $$ S = \int d^{N+1} x \sqrt{-g} e^{-2 \phi} [ R+ 4 (D \phi)^2 ] \eqno\eq $$ where $\phi$ is the dilaton. It is convenient to use new fields $\varphi$ and $\lambda$ defined by $$ a (t) = e^{\lambda t} \eqno\eq $$ and $$ \varphi = 2 \phi - N \lambda \, . \eqno\eq $$ The action (8.44) has the duality symmetry $$ \lambda \rightarrow - \lambda, \> \varphi \rightarrow \varphi \, . \eqno\eq $$ The variational equations of motion derived from (8.44) for a homogeneous and isotropic model are$^{228, 229)}$ $$ \eqalign{& \dot \varphi^2 = e^\varphi E + N \dot \lambda^2 \cr & \ddot \lambda - \dot \varphi \dot \lambda = {1\over 2} e^\varphi P \cr & \ddot \varphi = {1\over 2} e^\varphi E + N \dot \lambda^2 \, , }\eqno\eq $$ where $P$ and $E$ are total pressure and energy, respectively. For a winding mode-dominated equation of state (and neglecting friction terms) the equation of motion for $\lambda (t)$ becomes $$ \ddot \lambda = - {1\over{2N}} e^\varphi E(\lambda) \, , \eqno\eq $$ which corresponds to motion in a confining potential. Hence, winding modes prevent the background toroidal dimensions from expanding. These considerations may be used to put forward the conjecture$^{224)}$ that string cosmology will single out three as the maximum number of spatial dimensions which can be large ($R \gg 1$ in Planck units). The argument proceeds as follows. Space can, starting from an initial state with $R \sim 1$ in all directions, only expand if thermal equilibrium is maintained, which in turn is only possible if the winding modes can annihilate. This can only happen in at most three spatial dimensions (in a higher number the probability for intersection of the world sheets of two strings is zero). In the critical dimension for strings, $N=3$, the evolution of a string gas has been studied extensively in the context of the cosmic string theory (see Chapter 6). The winding modes do, indeed, annihilate, leaving behind a string network with about one winding mode passing through each Hubble volume. Thus, in string cosmology only three spatial dimensions will become large whereas the others will be confined to Planck size by winding modes. \section{Summary} Planck scale physics may have many observational consequences and may help cosmologists solve some of the deep puzzles concerning the origin of inflation, the absence of space-time singularities and the dimensionality of space-time. A lot of work needs to be done before these issues are properly understood. I have outlined two ways to address some of these questions. The first investigation was based on classical physics and attempted to analyze what can be said about the origin of inflation and about singularities from an effective action approach to gravity. We constructed a class of higher derivative gravity actions without singular cosmological solutions (i.e., no singular homogeneous and isotropic solutions) and which automatically give rise to inflation. The second approach was an exploration of some of the cosmological consequences of target space duality in string theory. A nonsingular cosmological scenario was proposed which might even explain why only three-spatial dimensions are large. \medskip \chapter{Conclusions} Modern cosmology has led to the development of several theories of structure formation, most prominently theories based on inflation, and topological defect models. These new theories are all based on the union between particle physics and general relativity. The models of structure formation obey the usual causality principle of relativistic physics. All of the current theories of structure formation have their problems. Most importantly, they do not address the cosmological constant problem but rather, inasmuch as they make use of scalar matter fields, make the problem worse. The inflationary Universe scenario is still lacking a convincing realization. Present versions require very special scalar field potentials. Topological defect models, on the other hand, do not explain why the Universe is nearly homogeneous and spatially flat (however, they are consistent with a low $\Omega$ Universe). In my opinion, we should regard our current theories as toy models with which we work and from which we learn, but which will eventually be replaced by improved and more convincing theories. Nevertheless, our present theories are predictive. To a first approximation, they all predict a scale invariant spectrum of density perturbations and induced CMB anisotropies. Typically, the models contain one intrinsically free parameter (plus maybe a couple more parameters with which we can describe our ignorance of the detailed evolution of the models). The free parameter can be normalized from any one of several observables. It is remarkable that the different normalizations of the models are consistent (to a first approximation). This lets us entertain the hope that we are on the right track: structure formation proceeds via gravitational instability (Sections 4.3 and 4.4) with the seed perturbations being provided by a particle physics theory of the very early Universe. There is already a wealth of observational data which is fit quite well by our present toy models. More and higher accuracy data is rapidly becoming available. The data concerns on one hand structure in the Universe gleamed from optical and infrared galaxy surveys, and on the other hand from the temperature map of the CMB sky. With the wealth of data available and steady flow of new observational results, and given that many important questions remain unresolved, modern cosmology will remain an exciting area of research for the forseeable future. The basic problems which are not addressed by our present theories of cosmology might be resolved by some as yet unknown unified theory of all forces. Some speculations along these lines were entertained in the last chapter of these lecture notes. It is of particular interest to investigate whether string theory leads to a more convincing realization of inflation, and whether there is a mechanism which predicts why our Universe consists of three large spatial dimensions. \centerline{\bf Acknowledgments} I wish to thank Professor Mario Novello for inviting me to give these lectures in Angra, and all the organizers and participants for their wonderful hospitality and for their many stimulating questions. I am grateful to all of my research collaborators, on whose work I have freely drawn. Partial financial support for the preparation of this manuscript has been provided at Brown by the US Department of Energy under Grant DE-FG0291ER40688, Task A, and at UBC by the Canadian NSERC under Grant 580441. \bigskip \REF\one{A. Linde, `Particle Physics and Inflationary Cosmology' (Harwood, Chur, 1990).} \REF\two{S. Blau and A. Guth, `Inflationary Cosmology,' in `300 Years of Gravitation' ed. by S. Hawking and W. Israel (Cambridge Univ. Press, Cambridge, 1987).} \REF\three{K. Olive, {\it Phys. Rep.} {\bf 190}, 307 (1990).} \REF\four{T.W.G. Kibble, {\it Phys. Rep.} {\bf 67}, 183 (1980).} \REF\five{A. Vilenkin, {\it Phys. Rep.} {\bf 121}, 263 (1985).} \REF\six{N. Turok, `Phase Transitions as the Origin of Large-Scale Structure,' in `Particles, Strings and Supernovae' (TASI-88) ed. by A. Jevicki and C.-I. Tan (World Scientific, Singapore, 1989).} \REF\seven{A. Vilenkin and E.P.S. Shellard, `Strings and Other Topological Defects' (Cambridge Univ. Press, Cambridge, 1994).} \REF\eight{R. Brandenberger, {\it Rev. Mod. Phys.} {\bf 57}, 1 (1985).} \REF\nine{R. Brandenberger, ``Modern Cosmology and Structure Formation", in `CP Violation and the Limits of the Standard Model (TASI94)', ed. J. Donoghue (World Scientific, Singapore, 1995).} \REF\ten{R. Brandenberger, in `Physics of the Early Universe,' proc. of the 1989 Scottish Univ. Summer School in Physics, ed. by J. Peacock, A. Heavens and A. Davies (SUSSP Publ., Edinburgh, 1990); \nextline R. Brandenberger, in `1991 Summer School in High Energy Physics and Cosmology', eds. E. Gava et al. (World Scientific, Singapore, 1992); \nextline R. Brandenberger, `Lectures on Modern Cosmology and Structure Formation', in `Particles and Fields', ed. by O. Eboli and V. Ribelles (World Scientific, Singapore 1994).} \REF\eleven{V. Mukhanov, H. Feldman and R. Brandenberger, {\it Phys. Rep.} {\bf 215}, 203 (1992).} \REF\twelve{ T.W.B. Kibble, {\it J. Phys.} {\bf A9}, 1387 (1976).} \REF\thirteen{E. Milne, {\it Zeits. f. Astrophys.} {\bf 6}, 1 (1933).} \REF\fourteen{V. de Lapparent, M. Geller and J. Huchra, {\it Ap. J. (Lett)} {\bf 302}, L1 (1986).} \REF\fifteen{S. Shechtman, P. Schechter, A. Oemler, D. Tucker, R. Kirshner and H. Lin, Harvard-Smithsonian preprint CFA 3385 (1992), to appear in `Clusters and Superclusters of Galaxies', ed. by A. Fabian (Kluwer, Dordrecht, 1993); \nextline S. Shechtman et al., `The Las Campanas Fiber-Optic Redshift Survey', CFA preprint (1994), to be publ. in the proc. of the 35th Herstmonceaux Conference `Wide Field Spectroscopy and the Distant Universe'.} \REF\sixteen{R. Partridge, {\it Rep. Prog. Phys.} {\bf 51}, 647 (1988).} \REF\seventeen{see e.g., S. Weinberg, `Gravitation and Cosmology' (Wiley, New York, 1972); \nextline Ya.B. Zel'dovich and I. Novikov, `The Structure and Evolution of the Universe' (Univ. of Chicago Press, Chicago, 1983).} \REF\eighteen{E. Hubble, {\it Proc. Nat. Acad. Sci.} {\bf 15}, 168 (1927).} \REF\nineteen{J. Mould et al., {\it Ap. J.} {\bf 383}, 467 (1991).} \REF\twenty{R. Alpher and R. Herman, {\it Rev. Mod. Phys.} {\bf 22}, 153 (1950); \nextline G. Gamov, {\it Phys. Rev.} {\bf 70}, 572 (1946).} \REF\twentyone{R. Dicke, P.J.E. Peebles, P. Roll and D. Wilkinson, {\it Ap. J.} {\bf 142}, 414 (1965).} \REF\twentytwo{A. Penzias and R. Wilson, {\it Ap. J.} {\bf 142}, 419 (1965).} \REF\twentythree{J. Mather et al., {\it Ap. J. (Lett.)} {\bf 354}, L37 (1990).} \REF\twentyfour{H. Gush, M. Halpern and E. Wishnow, {\it Phys. Rev. Lett.} {\bf 65}, 937 (1990).} \REF\twentyfive{R. Alpher, H. Bethe and G. Gamov, {\it Phys. Rev.} {\bf 73}, 803 (1948); \nextline R. Alpher and R. Herman, {\it Nature} {\bf 162}, 774 (1948).} \REF\twentysix{For an excellent introduction see S. Weinberg, `The First Three Minutes' (Basic Books, New York, 1988).} \REF\twseven{T. Padmanabhan, `Structure Formation in the Universe' (Cambridge Univ. Press, Cambridge, 1993), Chap. 3 and refs. therein.} \REF\tweight{H. Arp, G. Burbidge, F. Hoyle, J. Narlikar and N. Vickramasinghe, {\it Nature} {\bf 346}, 807 (1990).} \REF\twnine{P.J.E. Peebles, D. Schramm, E. Turner and R. Kron, {\it Nature} {\bf 352}, 769 (1991).} \REF\thirty{A. Guth, {\it Phys. Rev.} {\bf D23}, 347 (1981).} \REF\thone{P.J.E. Peebles, `Principles of Physical Cosmology' (Princeton Univ. Press, Princeton, 1993).} \REF\thtwo{J. Ostriker and L. Cowie, {\it Ap. J. (Lett.)} {\bf 243}, L127 (1981).} \REF\ththree{S. Weinberg, {\it Rev. Mod. Phys.} {\bf 61}, 1 (1989);\nextline S. Carroll, W. Press and E. Turner, {\it Ann. Rev. Astron. Astrophys.} {\bf 30}, 499 (1992).} \REF\thfour{D. Kazanas, {\it Ap. J.} {\bf 241}, L59 (1980).} \REF\thfive{W. Press, {\it Phys. Scr.} {\bf 21}, 702 (1980).} \REF\thsix{G. Chibisov and V. Mukhanov, `Galaxy Formation and Phonons,' Lebedev Physical Institute Preprint No. 162 (1980); \nextline G. Chibisov and V. Mukhanov, {\it Mon. Not. R. Astron. Soc.} {\bf 200}, 535 (1982).} \REF\thseven{V. Lukash, {\it Pis'ma Zh. Eksp. Teor. Fiz.} {\bf 31}, 631 (1980).} \REF\theight{K. Sato, {\it Mon. Not. R. Astron. Soc.} {\bf 195}, 467 (1981).} \REF\thnine{A. Starobinsky, {\it Phys. Lett.} {\bf 91B}, 99 (1980).} \REF\fourty{M. Mijic, M. Morris and W.-M. Suen, {\it Phys. Rev.} {\bf D34}, 2934 (1986).} \REF\foone{V. Mukhanov and R. Brandenberger, {\it Phys. Rev. Lett.} {\bf 68}, 1969 (1992); \nextline R. Brandenberger, V. Mukhanov and A. Sornborger, {\it Phys. Rev.} {\bf D48}, 1629 (1993).} \REF\fotwo{J. Ye and R. Brandenberger, {\it Nucl. Phys.} {\bf B346}, 149 (1990).} \REF\fothree{H. Nielsen and P. Olesen, {\it Nucl. Phys.} {\bf B61}, 45 (1973).} \REF\fofour{R. Davis, {\it Phys. Rev.} {\bf D35}, 3705 (1987).} \REF\fofive{N. Turok, {\it Phys. Rev. Lett.} {\bf 63}, 2625 (1989).} \REF\fosix{see e.g., G. Ross, Grand Unified Theories (Benjamin, Reading, 1985).} \REF\foseven{Ya.B. Zel'dovich, {\it Mon. Not. R. astron. Soc.} {\bf 192}, 663 (1980).} \REF\foeight{A. Vilenkin, {\it Phys. Rev. Lett.} {\bf 46}, 1169 (1981).} \REF\fonine{J. Primack, D. Seckel and B. Sadoulet, {\it Ann. Rev. Nucl. Part. Sci.} {\bf 38}, 751 (1988).} \REF\fifty{T. van Albada and R. Sancisi, {\it Phil. Trans. R. Soc. London}, {\bf A320}, 447 (1986).} \REF\fone{V. Trimble, {\it Ann. Rev. Astr. Astrophys.} {\bf 25}, 423 (1987).} \REF\ftwo{E. Bertschinger and A. Dekel, {\it Ap. J.} {\bf 336}, L5 (1989).} \REF\fthree{M. Strauss, M. Davis, A. Yahil and J. Huchra, {\it Ap. J.} {\bf 385}, 421 (1992).} \REF\ffour{S. White, C. Frenk and M. Davis, {\it Ap. J. (Lett.)} {\bf 274}, L1 (1993).} \REF\ffive{J. Bond, G. Efstathiou and J. Silk, {\it Phys. Rev. Lett.} {\bf 45}, 1980 (1980); \nextline J. Bond and A. Szalay, {\it Ap. J.} {\bf 274}, 443 (1983).} \REF\fsix{G. Blumenthal, S. Faber, J. Primack and M. Rees, {\it Nature} {\bf 311}, 517 (1984); \nextline M. Davis, G. Efstathiou, C. Frenk and S. White, {\it Ap. J.} {\bf 292}, 371 (1985).} \REF\fseven{N. Turok and R. Brandenberger, {\it Phys. Rev.} {\bf D33}, 2175 (1986); \nextline A. Stebbins, {\it Ap. J. (Lett.)} {\bf 303}, L21 (1986); \nextline H. Sato, {\it Prog. Theor. Phys.} {\bf 75}, 1342 (1986).} \REF\feight{T. Vachaspati, {\it Phys. Rev. Lett.} {\bf 57}, 1655 (1986).} \REF\fnine{A. Stebbins, S. Veeraraghavan, R. Brandenberger, J. Silk and N. Turok, {\it Ap. J.} {\bf 322}, 1 (1987).} \REF\sixty{N. Turok, {\it Phys. Scripta} {\bf T36}, 135 (1991).} \REF\sione{R. Brandenberger, N. Kaiser, D. Schramm and N. Turok, {\it Phys. Rev. Lett.} {\bf 59}, 2371 (1987).} \REF\sitwo{R. Brandenberger, N. Kaiser and N. Turok, {\it Phys. Rev.} {\bf D36}, 2242 (1987).} \REF\sithree{R. Brandenberger, L. Perivolaropoulos and A. Stebbins, {\it Int. J. of Mod. Phys.} {\bf A5}, 1633 (1990); \nextline L. Perivolarapoulos, R. Brandenberger and A. Stebbins, {\it Phys. Rev.} {\bf D41}, 1764 (1990); \nextline R. Brandenberger, {\it Phys. Scripta} {\bf T36}, 114 (1991).} \REF\sifour{M. Davis and J. Huchra, {\it Ap. J.} {\bf 254}, 437 (1982).} \REF\sifive{N. Bahcall and R. Soneira, {\it Ap. J.} {\bf 270}, 20 (1983); \nextline A. Klypin and A. Kopylov, {\it Sov. Astr. Lett.} {\bf 9}, 41 (1983).} \REF\sisix{M. Strauss et al., {\it Ap. J.} {\bf 385}, 421 (1992).} \REF\siseven{A. Yahil et al., {\it Ap. J.} {\bf 372}, 380 (1991).} \REF\sieight{G. Abell, {\it Ap. J. Suppl.} {\bf 3}, 211 (1958).} \REF\sinine{G. Daulton et al., `The two-point correlation function of rich clusters of galaxies: results from an extended APM cluster redshift survey', Oxford Univ. preprint, 1994, MNRAS, in press.} \REF\seventy{W. Zurek, {\it Ap. J.} {\bf 324}, 19 (1988).} \REF\sone{M. Rees and J. Ostriker, {\it Mon. Not. R. astr. Soc.} {\bf 179}, 541 (1977).} \REF\stwo{Ya.B. Zel'dovich, J. Einasto and S. Shandarin, {\it Nature} {\bf 300}, 407 (1982); \nextline J. Oort, {\it Ann. Rev. Astron. Astrophys.} {\bf 21}, 373 (1983); \nextline R.B. Tully, {\it Ap. J.} {\bf 257}, 389 (1982); \nextline S. Gregory, L. Thomson and W. Tifft, {\it Ap. J.} {\bf 243}, 411 (1980).} \REF\sthree{G. Chincarini and H. Rood, {\it Nature} {\bf 257}, 294 (1975); \nextline J. Einasto, M. Joeveer and E. Saar, {\it Mon. Not. R. astron. Soc.} {\bf 193}, 353 (1980); \nextline R. Giovanelli and M. Haynes, {\it Astron. J.} {\bf 87}, 1355 (1982); \nextline D. Batuski and J. Burns, {\it Ap. J.} {\bf 299}, 5 (1985).} \REF\sfour{R. Kirshner, A. Oemler, P. Schechter and S. Shechtman, {\it Ap. J. (Lett.)} {\bf 248}, L57 (1981).} \REF\sfive{M. Joeveer, J. Einasto and E. Tago, {\it Mon. Not. R. astron. Soc.} {\bf 185}, 357 (1978); \nextline L. da Costa et al., {\it Ap. J.} {\bf 327}, 544 (1988).} \REF\ssix{see e.g., G. Efstathiou, in `Physics of the Early Universe,' proc. of the 1989 Scottish Univ. Summer School in Physics, ed. by J. Peacock, A. Heavens and A. Davies (SUSSP Publ., Edinburgh, 1990).} \REF\sseven{T. Padmanabhan, `Structure Formation in the Universe' (Cambridge Univ. Press, Cambridge, 1993).} \REF\seight{E. Harrison, {\it Phys. Rev.} {\bf D1}, 2726 (1970); \nextline Ya.B. Zel'dovich, {\it Mon. Not. R. astron. Soc.} {\bf 160}, 1p (1972).} \REF\snine{E. Lifshitz, {\it Zh. Eksp. Teor. Fiz.} {\bf 16}, 587 (1946); \nextline E. Lifshitz and I. Khalatnikov, {\it Adv. Phys.} {\bf 12}, 185 (1963).} \REF\eighty{W. Press and E. Vishniac, {\it Ap. J.} {\bf 239}, 1 (1980).} \REF\eone{R. Brandenberger, H. Feldman, V. Mukhanov and T. Prokopec, `Gauge Invariant Cosmological Perturbations: Theory and Applications,' publ. in ``The Origin of Structure in the Universe," eds. E. Gunzig and P. Nardone (Kluwer, Dordrecht, 1993).} \REF\etwo{J. Bardeen, {\it Phys. Rev.} {\bf D22}, 1882 (1980).} \REF\ethree{R. Brandenberger, R. Kahn and W. Press, {\it Phys. Rev.} {\bf D28}, 1809 (1983).} \REF\efour{H. Kodama and M. Sasaki, {\it Prog. Theor. Phys. Suppl.} No. 78, 1 (1984).} \REF\efive{R. Durrer and N. Straumann, {\it Helvet. Phys. Acta} {\bf 61}, 1027 (1988).} \REF\esix{D. Lyth and M. Mukherjee, {\it Phys. Rev.} {\bf D38}, 485 (1988).} \REF\eseven{G.F.R. Ellis and M. Bruni, {\it Phys. Rev.} {\bf D40}, 1804 (1989).} \REF\eeight{J. Stewart, {\it Class. Quantum Grav.} {\bf 7}, 1169 (1990).} \REF\enine{J. Stewart and M. Walker, {\it Proc. R. Soc.} {\bf A341}, 49 (1974).} \REF\ninety{J. Bardeen, P. Steinhardt and M. Turner, {\it Phys. Rev.} {\bf D28}, 679 (1983).} \REF\none{D. Kirzhnits and A. Linde, {\it Pis'ma Zh. Eksp. Teor. Fiz.} {\bf 15}, 745 (1972); \nextline D. Kirzhnits and A. Linde, {\it Zh. Eksp. Teor. Fiz.} {\bf 67}, 1263 (1974);\nextline C. Bernard, {\it Phys. Rev.} {\bf D9}, 3313 (1974);\nextline L. Dolan and R. Jackiw, {\it Phys. Rev.} {\bf D9}, 3320 (1974);\nextline S. Weinberg, {\it Phys. Rev.} {\bf D9}, 3357 (1974).} \REF\ntwo{A. Linde, {\it Phys. Lett.} {\bf 129B}, 177 (1983).} \REF\nthree{G. Mazenko, W. Unruh and R. Wald, {\it Phys. Rev.} {\bf D31}, 273 (1985).} \REF\nfour{A. Linde, {\it Phys. Lett.} {\bf 108B}, 389 (1982); \nextline A. Albrecht and P. Steinhardt, {\it Phys. Rev. Lett.} {\bf 48}, 1220 (1982).} \REF\nfive{J. Langer, {\it Physica} {\bf 73}, 61 (1974).} \REF\nsix{S. Coleman, {\it Phys. Rev.} {\bf D15}, 2929 (1977); \nextline C. Callan and S. Coleman, {\it Phys. Rev.} {\bf D16}, 1762 (1977).} \REF\nseven{M. Voloshin, Yu. Kobzarev and L. Okun, {\it Sov. J. Nucl. Phys.} {\bf 20}, 644 (1975).} \Ref\neight{M. Stone, {\it Phys. Rev.} {\bf D14}, 3568 (1976);\nextline M. Stone, {\it Phys. Lett.} {\bf 67B}, 186 (1977).} \Ref\nnine{P. Frampton, {\it Phys. Rev. Lett.}, {\bf 37}, 1380 (1976).} \Ref\hundred{S. Coleman, in `The Whys of Subnuclear Physics' (Erice 1977), ed by A. Zichichi (Plenum, New York, 1979).} \REF\hone{A. Guth and S.-H. Tye, {\it Phys. Rev. Lett.} {\bf 44}, 631 (1980).} \Ref\htwo{A. Guth and E. Weinberg, {\it Nucl. Phys.} {\bf B212}, 321 (1983).} \Ref\hthree{S. Hawking and I. Moss, {\it Phys. Lett.} {\bf 110B}, 35 (1982).} \Ref\hfour{R. Matzner, in `Proceedings of the Drexel Workshop on Numerical Relativity', ed by J. Centrella (Cambridge Univ. Press, Cambridge, 1986).} \Ref\hfive{A. Albrecht, P. Steinhardt, M. Turner and F. Wilczek, {\it Phys. Rev. Lett.} {\bf 48}, 1437 (1982).} \Ref\hsix{L. Abbott, E. Farhi and M. Wise, {\it Phys. Lett.} {\bf 117B}, 29 (1982).} \Ref\hseven{J. Traschen and R. Brandenberger, {\it Phys. Rev.} {\bf D42}, 2491 (1990).} \REF\height{L. Kofman, A. Linde and A. Starobinski, {\it Phys. Rev. Lett.} {\bf 73}, 3195 (1994);\nextline Y. Shtanov, J. Traschen and R. Brandenberger, {\it Phys. Rev.} {\bf D51}, 5438 (1995).} \REF\hnine{L. Landau and E. Lifshitz, `Mechanics' (Pergamon, Oxford, 1960); \nextline V. Arnold, `Mathematical Methods of Classical Mechanics' (Springer, New York, 1978).} \REF\hten{S. Coleman and E. Weinberg, {\it Phys. Rev.} {\bf D7}, 1888 (1973).} \REF\htone{M. Markov and V. Mukhanov, {\it Phys. Lett.} {\bf 104A}, 200 (1984); \nextline V. Belinsky, L. Grishchuk, I. Khalatnikov and Ya. Zel'dovich, {\it Phys. Lett.} {\bf 155B}, 232 (1985); \nextline L. Kofman, A. Linde and A. Starobinsky, {\it Phys. Lett.} {\bf 157B}, 36 (1985); \nextline T. Piran and R. Williams, {\it Phys. Lett.} {\bf 163B}, 331 (1985).} \REF\httwo{S.-Y. Pi, {\it Phys. Rev. Lett.} {\bf 52}, 1725 (1984); \nextline K. Freese, J. Frieman and A. Olinto, {\it Phys. Rev. Lett.} {\bf 65}, 3233 (1990).} \REF\htthree{D. Nanopoulos, K. Olive, M. Srednicki and K. Tamvakis, {\it Phys. Lett.} {\bf 123B}, 41 (1983); \nextline J. Ellis, K. Enqvist, D. Nanopoulos, K. Olive and M. Srednicki, {\it Phys. Lett.} {\bf 152B}, 175 (1985); \nextline R. Holman, P. Ramond and C. Ross, {\it Phys. Lett.} {\bf 137B}, 343 (1984).} \REF\htfour{A. Goncharov and A. Linde, {\it JETP} {\bf 59}, 930 (1984); \nextline A. Goncharov and A. Linde, {\it Phys. Lett.} {\bf 139B}, 27 (1984); \nextline A. Goncharov and A. Linde, {\it Class. Quant. Grav.} {\bf 1}, L75 (1984).} \REF\htfive{I. Antoniadis, J. Ellis, J. Hagelin and D. Nanopoulos, {\it Phys. Lett.} {\bf 205B}, 459 (1988); \nextline I. Antoniadis, J. Ellis, J. Hagelin and D. Nanopoulos, {\it Phys. Lett.} {\bf 208B}, 209 (1988).} \REF\htsix{A. Linde, D. Linde and A. Mezhlumian, {\it Phys. Rev.} {\bf D49}, 1783 (1994); \nextline A. Linde, `Lectures on Inflationary Cosmology', Stanford preprint SU-ITP-94-36, hep-th/9410082 (1994).} \REF\htseven{A. Starobinsky, in `Current Trends in Field Theory, Quantum Gravity, and Strings', Lecture Notes in Physics, ed. by H. de Vega and N. Sanchez (Springer, Heidelberg, 1986).} \REF\hteight{B. Whitt, {\it Phys. Lett.} {\bf 145B}, 176 (1984).} \REF\htnine{M. Gasperini and G. Veneziano, {\it Astropart. Phys.} {\bf 1}, 317 (1993); \nextline R. Brustein and G. Veneziano, {\it Phys. Lett.} {\bf B329}, 429 (1994).} \REF\htw{F. Adams, K. Freese and A. Guth, {\it Phys. Rev.} {\bf D43}, 965 (1991).} \REF\htwone{H. Feldman and R. Brandenberger, {\it Phys. Lett.} {\bf 227B}, 359 (1989).} \REF\htwtwo{J. Kung and R. Brandenberger, {\it Phys. Rev.} {\bf D42}, 1008 (1990).} \REF\htwthree{D. Goldwirth and T. Piran, {\it Phys. Rev. Lett.} {\bf 64}, 2852 (1990);\nextline D. Goldwirth and T. Piran, {\it Phys. Rep.} {\bf 214}, 223 (1992).} \REF\htwfour{A. Albrecht and R. Brandenberger, {\it Phys. Rev.} {\bf D31}, 1225 (1985).} \REF\htwfive{G. Chibisov and V. Mukhanov, {\it JETP Lett.} {\bf 33}, 532 (1981);\nextline V. Mukhanov and G. Chibisov, {\it Zh. Eksp. Teor. Fiz.} {\bf 83}, 475 (1982).} \REF\htwsix{A. Lapedes, {\it J. Math. Phys.} {\bf 19}, 2289 (1978); \nextline R. Brandenberger and R. Kahn, {\it Phys. Lett.} {\bf 119B}, 75 (1982).} \REF\htwseven{A. Guth and S.-Y. Pi, {\it Phys. Rev. Lett.} {\bf 49}, 1110 (1982); \nextline S. Hawking, {\it Phys. Lett.} {\bf 115B}, 295 (1982); \nextline A. Starobinsky, {\it Phys. Lett.} {\bf 117B}, 175 (1982); \nextline R. Brandenberger and R. Kahn, {\it Phys. Rev.} {\bf D28}, 2172 (1984); \nextline J. Frieman and M. Turner, {\it Phys. Rev.} {\bf D30}, 265 (1984); \nextline V. Mukhanov, {\it JETP Lett.} {\bf 41}, 493 (1985).} \REF\htweight{W. Zurek, {\it Phys. Rev.} {\bf D24}, 1516 (1982); \nextline W. Zurek, {\it Phys. Rev.} {\bf D26} 1862 (1982).} \REF\htwnine{E. Joos and H. Zeh, {\it Z. Phys.} {\bf B59}, 223 (1985); \nextline H. Zeh, {\it Phys. Lett.} {\bf 116A}, 9 (1986); \nextline C. Kiefer, {\it Class. Quantum Grav.} {\bf 4}, 1369 (1987); \nextline T. Fukuyama and M. Morikawa, {\it Phys. Rev.} {\bf D39}, 462 (1989); \nextline J. Halliwell, {\it Phys. Rev.} {\bf D39}, 2912 (1989); \nextline T. Padmanabhan, {\it Phys. Rev.} {\bf D39}, 2924 (1980); \nextline W. Unruh and W. Zurek, {\it Phys. Rev.} {\bf D40}, 1071 (1989); \nextline E. Calzetta and F. Mazzitelli, {\it Phys. Rev.} {\bf D42}, 4066 (1990); \nextline S. Habib and R. Laflamme, {\it Phys. Rev.} {\bf D42}, 4056 (1990); \nextline H. Feldman and A. Kamenshchik, {\it Class. Quantum Grav.} {\bf 8}, L65 (1991).} \REF\hthirty{M. Sakagami, {\it Prog. Theor. Phys.} {\bf 79}, 443 (1988);\nextline R. Brandenberger, R. Laflamme and M. Mijic, {\it Mod. Phys. Lett.} {\bf A5}, 2311 (1990).} \REF\hthone{J. Bardeen, unpublished (1984).} \REF\hthtwo{R. Brandenberger, {\it Nucl. Phys.} {\bf B245}, 328 (1984).} \REF\hththree{N. Birrell and P. Davies, `Quantum Fields in Curved Space' (Cambridge Univ. Press, Cambridge, 1982).} \REF\hthfour{R. Brandenberger and C. Hill, {\it Phys. Lett.} {\bf 179B}, 30 (1986).} \REF\hthfive{W. Fischler, B. Ratra and L. Susskind, {\it Nucl. Phys.} {\bf B259}, 730 (1985).} \REF\hthsix{D. Mermin, {\it Rev. Mod. Phys.} {\bf 51}, 591 (1979).} \REF\hthseven{P. de Gennes, `The Physics of Liquid Crystals' (Clarendon Press, Oxford, 1974); \nextline I. Chuang, R. Durrer, N. Turok and B. Yurke, {\it Science} {\bf 251}, 1336 (1991); \nextline M. Bowick, L. Chandar, E. Schiff and A. Srivastava, {\it Science} {\bf 263}, 943 (1994).} \REF\htheight{M. Salomaa and G. Volovik, {\it Rev. Mod. Phys.} {\bf 59}, 533 (1987).} \REF\hthnine{A. Abrikosov, {\it JETP} {\bf 5}, 1174 (1957).} \REF\hfourty{Ya.B. Zel'dovich, I. Kobzarev and L. Okun, {\it Zh. Eksp. Teor. Fiz.} {\bf 67}, 3 (1974).} \REF\hfoone{Ya.B. Zel'dovich and M. Khlopov, {\it Phys. Lett.} {\bf 79B}, 239 (1978); \nextline J. Preskill, {\it Phys. Rev. Lett.} {\bf 43}, 1365 (1979).} \REF\hfotwo{T.W.B. Kibble, {\it Acta Physica Polonica} {\bf B13}, 723 (1982).} \REF\hfothree{P. Langacker and S.-Y. Pi, {\it Phys. Rev. Lett.} {\bf 45}, 1 (1980).} \REF\hfofour{T.W.B. Kibble and E. Weinberg, {\it Phys. Rev.} {\bf D43}, 3188 (1991).} \REF\hfoseven{T. Vachaspati and A. Vilenkin, {\it Phys. Rev.} {\bf D30}, 2036 (1984).} \REF\hfoeight{S. Rudaz and A. Srivastava, {\it Mod. Phys. Lett.} {\bf A8}, 1443 (1993).} \REF\hfnine{F. Liu, M. Mondello and N. Goldenfeld, {\it Phys. Rev. Lett.} {\bf 66}, 3071 (1991).} \REF\hsixty{M. Hindmarsh, A.-C. Davis and R. Brandenberger, {\it Phys. Rev.} {\bf D49}, 1944 (1994); \nextline R. Brandenberger and A.-C. Davis, {\it Phys. Lett.} {\bf B332}, 305 (1994).} \REF\hsione{T. Prokopec, A. Sornborger and R. Brandenberger, {\it Phys. Rev.} {\bf D45}, 1971 (1992).} \REF\hsitwo{J. Borrill, E. Copeland and A. Liddle, {\it Phys. Lett.} {\bf 258B}, 310 (1991).} \REF\hsithree{A. Sornborger, {\it Phys. Rev.} {\bf D48}, 3517 (1993).} \REF\hsifour{L. Perivolaropoulos, {\it Phys. Rev.} {\bf D46}, 1858 (1992).} \REF\hsifive{T. Prokopec, {\it Phys. Lett.} {\bf 262B}, 215 (1991);\nextline R. Leese and T. Prokopec, {\it Phys. Rev.} {\bf D44}, 3749 (1991).} \REF\hfofive{M. Barriola and A. Vilenkin, {\it Phys. Rev. Lett.} {\bf 63}, 341 (1989).} \REF\hfosix{S. Rhie and D. Bennett, {\it Phys. Rev. Lett.} {\bf 65}, 1709 (1990).} \REF\hsisix{D. Foerster, {\it Nucl. Phys.} {\bf B81}, 84 (1974).} \REF\hsiseven{N. Turok, in `Proceedings of the 1987 CERN/ESO Winter School on Cosmology and Particle Physics' (World Scientific, Singapore, 1988).} \REF\hsieight{T.W.B. Kibble and N. Turok, {\it Phys. Lett.} {\bf 116B}, 141 (1982).} \REF\hsinine{R. Brandenberger, {\it Nucl. Phys.} {\bf B293}, 812 (1987).} \REF\hseventy{E.P.S. Shellard, {\it Nucl. Phys.} {\bf B283}, 624 (1987).} \REF\hsone{R. Matzner, {\it Computers in Physics} {\bf 1}, 51 (1988); \nextline K. Moriarty, E. Myers and C. Rebbi, {\it Phys. Lett.} {\bf 207B}, 411 (1988); \nextline E.P.S. Shellard and P. Ruback, {\it Phys. Lett.} {\bf 209B}, 262 (1988).} \REF\hstwo{P. Ruback, {\it Nucl. Phys.} {\bf B296}, 669 (1988).} \REF\hsthree{A. Albrecht and N. Turok, {\it Phys. Rev. Lett.} {\bf 54}, 1868 (1985).} \REF\hsfour{D. Bennett and F. Bouchet, {\it Phys. Rev. Lett.} {\bf 60}, 257 (1988).} \REF\hsfive{B. Allen and E.P.S. Shellard, {\it Phys. Rev. Lett.} {\bf 64}, 119 (1990).} \REF\hssix{A. Albrecht and N. Turok, {\it Phys. Rev. } {\bf D40}, 973 (1989).} \REF\hsseven{R. Brandenberger and J. Kung, in `The Formation and Evolution of Cosmic Strings' eds. G. Gibbons, S. Hawking and T. Vachaspati (Cambridge Univ. Press, Cambridge, 1990).} \REF\hseight{see e.g., C. Misner, K. Thorne and J. Wheeler, `Gravitation' (Freeman, San Francisco, 1973).} \REF\hsnine{T. Vachaspati and A. Vilenkin, {\it Phys. Rev.} {\bf D31}, 3052 (1985); \nextline N. Turok, {\it Nucl. Phys.} {\bf B242}, 520 (1984); \nextline C. Burden, {\it Phys. Lett.} {\bf 164B}, 277 (1985).} \REF\heighty{B. Carter, {\it Phys. Rev.} {\bf D41}, 3869 (1990).} \REF\heone{E. Copeland, T.W.B. Kibble and D. Austin, {\it Phys. Rev.} {\bf D45}, 1000 (1992).} \REF\hetwo{N. Turok, {\it Nucl. Phys.} {\bf B242}, 520 (1984).} \REF\hethree{J. Silk and V. Vilenkin, {\it Phys. Rev. Lett.} {\bf 53}, 1700 (1984).} \REF\hefour{A. Vilenkin, {\it Phys. Rev.} {\bf D23}, 852 (1981); \nextline J. Gott, {\it Ap. J.} {\bf 288}, 422 (1985); \nextline W. Hiscock, {\it Phys. Rev.} {\bf D31}, 3288 (1985); \nextline B. Linet, {\it Gen. Rel. Grav.} {\bf 17}, 1109 (1985); \nextline D. Garfinkle, {\it Phys. Rev.} {\bf D32}, 1323 (1985); \nextline R. Gregory, {\it Phys. Rev. Lett.} {\bf 59}, 740 (1987).} \REF\hefive{D. Vollick, {\it Phys. Rev.} {\bf D45}, 1884 (1992); \nextline T. Vachaspati and A. Vilenkin, {\it Phys. Rev. Lett.} {\bf 67}, 1057 (1991).} \REF\hesix{A. Albrecht and A. Stebbins, {\it Phys. Rev. Lett.} {\bf 68}, 2121 (1992).} \REF\heseven{A. Albrecht and A. Stebbins, {\it Phys. Rev. Lett.} {\bf 69}, 2615 (1992).} \REF\heeight{D. Spergel, N. Turok, W. Press and B. Ryden, {\it Phys. Rev.} {\bf D43}, 1038 (1991).} \REF\henine{A. Gooding, D. Spergel and N. Turok, {\it Ap. J. (Lett.)} {\bf 372}, L5 (1991); \nextline C. Park, D. Spergel and N. Turok, {\it Ap. J. (Lett.)} {\bf 373}, L53 (1991).} \REF\hninety{R. Cen, J. Ostriker, D. Spergel and N. Turok, {\it Ap. J.} {\bf 383}, 1 (1991).} \REF\hnone{J. Gott, A. Melott and M. Dickinson, {\it Ap. J.} {\bf 306}, 341 (1986).} \REF\hntwo{S. Ramsey, Senior thesis, Brown Univ. (1992); \nextline D. Kaplan, Senior thesis, Brown Univ. (1993); \nextline R. Brandenberger, D. Kaplan and S. Ramsey, `Some statistics for measuring large-scale structure', Brown preprint BROWN-HET-922 (1993); \nextline A. Aguirre, Senior thesis, Brown Univ. (1995).} \REF\hnthree{D. Fixen, E. Cheng and D. Wilkinson, {\it Phys. Rev. Lett.} {\bf 50}, 620 (1983).} \REF\hnfour{R. Sachs and A. Wolfe, {\it Ap. J.} {\bf 147}, 73 (1967).} \REF\hnfive{V. Mukhanov and G. Chibisov, {\it Sov. Astron. Lett.} {\bf 10}, 890 (1984).} \REF\hnsix{N. Kaiser and A. Stebbins, {\it Nature} {\bf 310}, 391 (1984).} \REF\hnseven{R. Moessner, L. Perivolaropoulos and R. Brandenberger, {\it Ap. J.} {\bf 425}, 365 (1994).} \REF\hneight{N. Turok and D. Spergel, {\it Phys. Rev. Lett.} {\bf 64}, 2736 (1990).} \REF\thundred{R. Durrer, A. Howard and Z.-H. Zhou, {\it Phys. Rev.} {\bf D49}, 681 (1994).} \REF\twhone{U.-L. Pen, D. Spergel and N. Turok, {\it Phys. Rev.} {\bf D49}, 692 (1994).} \REF\twhtwo{E. Wright et al., {\it Ap. J. (Lett.)} {\bf 396}, L13 (1992).} \REF\twohthree{S. White, C. Frenk, M. Davis and G. Efstathiou, {\it Ap. J.} {\bf 313}, 505 (1987).} \REF\twohfive{D. Bennett and S. Rhie, {\it Ap. J. (Lett.)} {\bf 406}, L7 (1993).} \REF\twohfour{D. Bennett, A. Stebbins and F. Bouchet, {\it Ap. J. (Lett.)} {\bf 399}, L5 (1992).} \REF\twohsix{L. Perivolaropoulos, {\it Phys. Lett.} {\bf 298B}, 305 (1993).} \REF\twohseven{G. Smoot et al., {\it Ap. J. (Lett.)} {\bf 396}, L1 (1992).} \REF\twoheight{S. Meyer, E. Cheng and L. Page, {\it Ap. J. (Lett.)} {\bf 371}, L7 (1991);\nextline K. Ganga, E. Cheng, S. Meyer and L. Page, {\it Ap. J. (Lett.)} {\bf 410}, L57 (1993).} \REF\twohnine{S. Hancock et al., {\it Nature} {\bf 317}, 333 (1994).} \REF\twohten{E. Wollack et al., {\it Ap. J. (Lett.)} {\bf 419}, L49 (1993).} \REF\twoheleven{T. Gaier et al., {\it Ap. J. (Lett.)} {\bf 398}, L1 (1992); \nextline J. Schuster et al., {\it Ap. J. (Lett.)} {\bf 412}, L47 (1993).} \REF\twohtwelve{P. de Bernardis et al., {\it Ap. J. (Lett.)} {\bf 422}, L33 (1994).} \REF\twohthteen{M. Dragovan et al., Princeton preprint (1993).} \REF\twohfourteen{P. Meinhold et al., {\it Ap. J. (Lett.)} {\bf 409}, L1 (1993).} \REF\twohfifteen{J. Gunderson et al., {\it Ap. J. (Lett.)} {\bf 413}, L1 (1993).} \REF\twohsixteen{E. Cheng et al., {\it Ap. J. (Lett.)} {\bf 422}, L37 (1994).} \REF\twohseventeen{G. Tucker, G. Griffin, H. Nguyen and J. Peterson, {\it Ap. J. (Lett.)} {\bf 419}, L45 (1993).} \REF\twoheighteen{A. Readhead et al., {\it Ap. J.} {\bf 346}, 566 (1989).} \REF\twohnineteen{S. Myers, A. Readhead and A. Lawrence, {\it Ap. J.} {\bf 405}, 8 (1993).} \REF\twotw{W. Freedman et al., {\it Nature} {\bf 371}, 757 (1994); \nextline M. Pierce et al., {\it Nature} {\bf 371}, 385 (1994).} \REF\twotwone{A. Riess, W. Press and R. Kirshner, {\it Ap. J. (Lett.)} {\bf 438}, L17 (1995).} \REF\twotwtwo{M. Jones et al., {\it Nature} {\bf 365}, 320 (1993).} \REF\twotwthree{R. Brandenberger, {\it Phys. Lett.} {\bf B129}, 397 (1983).} \REF\twotwfour{N. Tsamis and R. Woodard, {\it Phys. Lett.} {\bf B301}, 351 (1993).} \REF\twotwfive{W. Zurek, {\it Acta Phys. Pol.} {\bf B24}, 1301 (1993).} \REF\twotwsix{T. Kibble and A. Vilenkin, ``Density of strings formed at a second order cosmological phase transition", Imperial preprint IMPERIAL-TP-94/95-9A, hep-ph/9501207 (1995).} \REF\twotwseven{A. Aguirre and R. Brandenberger, ``Accretion of hot dark matter onto slowly moving cosmic strings", Brown preprint BROWN-HET-995, astro-ph/9505031, {\it Int. J. Mod. Phys. D}, in press (1995).} \REF\twotweight{R. Moessner and R. Brandenberger, ``Formation of high redshift objects in a cosmic string model with hot dark matter", Brown preprint BROWN-HET-1001 (1995).} \REF\ffsix{M. Trodden, V. Mukhanov and R. Brandenberger, {\it Phys. Lett.} {\bf B316}, 483 (1993).} \REF\ffseven{R. Moessner and M. Trodden, {\it Phys. Rev.} {\bf D51}, 2801 (1995).} \REF\ffeight{B. Altshuler, {\it Class. Quant. Grav.} {\bf 7}, 189 (1990).} \REF\ffnine{M. Markov, {\it Pis'ma Zh. Eksp. Theor. Fiz.} {\bf 36}, 214 (1982); \nextline M. Markov, {\it Pis'ma Zh. Eksp. Theor. Fiz.} {\bf 46}, 342 (1987); \nextline V. Ginsburg, V. Mukhanov and V. Frolov, {\it Pis'ma Zh. Eksp. Theor. Fiz.} {\bf 94}, 3 (1988); \nextline V. Frolov, M. Markov and V. Mukhanov, {\it Phys. Rev.} {\bf D41}, 383 (1990).} \REF\fften{R. Brandenberger, M. Mohazzab, V. Mukhanov, A. Sornborger and M. Trodden, in preparation (1995).} \REF\fftwelve{K. Kikkawa and M. Yamasaki, {\it Phys. Lett.} {\bf B149}, 357 (1984); \nextline N. Sakai and I. Senda, {\it Prog. Theor. Phys.} {\bf 75}, 692 (1986); \nextline B. Sathiapalan, {\it Phys. Rev. Lett.} {\bf 58}, 1597 (1987); \nextline P. Ginsparg and C. Vafa, {\it Nucl. Phys.} {\bf B289}, 414 (1987).} \REF\ffthirteen{R. Brandenberger and C. Vafa, {\it Nucl. Phys.} {\bf B316}, 391 (1989).} \REF\fffourteen{R. Hagedorn, {\it Nuovo Cimento Suppl.} {\bf 3}, 147 (1965).} \REF\fffifteen{D. Mitchell and N. Turok, {\it Nucl. Phys.} {\bf B294}, 1138 (1987).} \REF\ffsixteen{N. Deo, S. Jain and C.-I. Tan, {\it Phys. Rev.} {\bf D40}, 2626 (1989).} \REF\ffseventeen{A. Tseytlin and C. Vafa, {\it Nucl. Phys.} {\bf B372}, 443 (1992).} \REF\ffeighteen{G. Veneziano, {\it Phys. Lett.} {\bf B265}, 287 (1991).} \refout \end
\section*{\large \bf References} \parskip=.5ex plus 1.0pt \def\bibitem{\par \noindent \hangindent\parindent \hangafter=1}} \def\par \endgroup{\par \endgroup} \begin{document} \begin{titlepage} \pagestyle{empty} \baselineskip=21pt \rightline{UMN-TH-1402/95} \rightline{astro-ph/9508086} \rightline{August 1995} \vskip .2in \begin{center} {\large{\bf The Local Abundance of $^3$He: A Confrontation Between Theory and Observation}} \end{center} \vskip .1in \begin{center} Sean T. Scully$^1$, Michel Cass\'e$^2$, Keith A. Olive$^1$, David N. Schramm$^{3,4}$, James Truran$^3$, and Elisabeth Vangioni-Flam$^5$ $^1${\it School of Physics and Astronomy, University of Minnesota} {\it Minneapolis, MN 55455, USA} $^2${\it Service d'Astrophysique, DSM, DAPNIA, CEA, France} $^3${\it Department of Astronomy and Astrophysics, Enrico Fermi Institute, The University of Chicago, Chicago, IL 60637-1433} $^4${\it NASA/Fermilab Astrophysics Center, Fermi National Accelerator Laboratory, Batavia, IL 60510-0500} $^5${\it Institut d'Astrophysique de Paris, 98bis Boulevard Arago, 75014 Paris, France} \end{center} \vskip .5in \centerline{ {\bf Abstract} } \baselineskip=18pt Determinations of the \he3 concentrations in Galactic matter serve to impose interesting and important constraints both on cosmological models and on models of Galactic chemical evolution. At present, observations of \he3 in the solar system and in the interstellar medium today suggest that the \he3 abundance has not increased significantly over the history of the Galaxy, while theoretical models of Galactic chemical evolution (utilizing current nucleosynthesis yields from stellar evolution and supernova models) predict a rather substantial increase in \he3. We consider the possibility that the solar \he3 abundance may have been affected by stellar processing in the solar neighborhood prior to the formation of the solar system. Such a discrepancy between solar abundances and average galactic abundances by as much as a factor of two, may be evidenced by several isotopic anomalies. Local destruction of \he3 by a similar amount could serve to help to reconcile the expected increase in the \he3 abundance predicted by models of galactic chemical evolution. We find however, that the production of heavier elements, such as oxygen, places a strong constraint on the degree of \he3 destruction. We also explore the implications of both alternative models of Galactic chemical evolution and the stellar yields for \he3 in low mass stars, which can explain the history of the \he3 concentration in the Galaxy. \noindent \end{titlepage} \baselineskip=18pt \def~\mbox{\raisebox{-.7ex}{$\stackrel{<}{\sim}$}}~{~\mbox{\raisebox{-.7ex}{$\stackrel{<}{\sim}$}}~} \def~\mbox{\raisebox{-.7ex}{$\stackrel{>}{\sim}$}}~{~\mbox{\raisebox{-.7ex}{$\stackrel{>}{\sim}$}}~} \def\begin{equation}{\begin{equation}} \def\end{equation}{\end{equation}} \section{Introduction} There is an inherent difficulty associated with the utilization of the observed abundances of D and \he3 to predict their primordial values. Namely, the connection between the primordial abundances of D and \he3 and their solar or present-day values depends sensitively on models of galactic chemical evolution. In principle, measurements of D in quasar absorption systems could dramatically help us to bridge this gap, by providing directly the primordial abundance of D and hence the value of the baryon-to-photon ratio, $\eta$, from big bang nucleosynthesis (Walker {\it et al.}~ 1991). However, recent measurements of this kind (Carswell {\it et al.}~ 1994; Songaila {\it et al.}~ 1994; Tytler \& Fan 1995) must be viewed as preliminary, as the determined D abundance in the two absorption systems observed are not concordant with each other. Until such measurements yield a single consistent value for primordial D, we must continue to be guided by models of galactic chemical evolution. It has been well established that models of galactic chemical evolution, consistent with the constraints imposed by element abundance determinations, are capable of destroying significant amounts of deuterium (Truran \& Cameron 1971; Gry {\it et al.}~ 1984; Delbourgo-Salvador {\it et al.}~ 1985; Vangioni-Flam and Audouze 1988; Vangioni-Flam, Olive \& Prantzos 1994; Vangioni-Flam \& Cass\'{e}, 1995). However, as was recently discussed in Olive {\it et al.}~ (1995), the problem rests not with the destruction of D, but rather with the production of \he3. Though \he3 is partially destroyed in massive stars, \he3 production in low mass stars generally leads to a net increase in the \he3 abundance over the evolutionary history of the galaxy. Observations of large \he3 enhancements in planetary nebulae (Rood, Bania \& Wilson, 1992; Rood, Bania, Wilson \& Balser, 1995) support the conclusion that \he3 is indeed produced in low mass stars. An excess of \he3 is difficult to avoid if low mass stars are strong producers of this isotope, as indicated by the calculations of Iben and Truran (1978) and more recently Vassiliadis and Wood (1993) and Weiss {\it et al.}~ (1995). However, recent models with increased mixing have been calculated, which both bring the carbon and oxygen isotopic ratios to their observed level in red giants and lead to a net destruction of \he3 (Charbonnel 1994; Wasserburg {\it et al.}~ 1995; Hogan 1995). At present, these models are quite preliminary, and it is premature to draw a firm conclusion. However, if their results are confirmed, \he3 will be significantly less problematic, unless the primordial D abundance is as high as observed by Songaila {\it et al.}~ (1994) and Carswell {\it et al.}~ (1994); in this case we would still require \he3 destruction factors in excess of what current calculations bear out (Olive {\it et al.}~ 1995). Abundance determinations of \he3 at the time of the formation of the solar system seem to indicate that the solar \he3 abundance is very close to that of the primordial abundance. Here, we will examine in some detail the possibility that the \he3 abundance in the solar system may be depleted with respect to galactic averages at the time of its formation. We find, however, that the abundances of the heavier elements, most notably oxygen and neon, impose a strong constraint on the degree of depletion of \he3. We are therefore left with the following possibilities: either the initial deuterium abundance is low, D/H $\sim 3 \times 10^{-5}$ (we will discuss rough lower limits imposed by models of chemical evolution); or more dramatic changes are required in models of chemical evolution, which have the effect of maintaining a rather flat evolution of \he3 with time (we will show an example of this type of model below); or the stellar yields of \he3 in low mass stars are lower than previously thought. Because deuterium is converted to \he3 in the pre-main-sequence phase of stellar evolution, models without a significantly depressed initial D abundance are subject to problems with \he3 production. If the initial abundance of D is rather low (D $ ~\mbox{\raisebox{-.7ex}{$\stackrel{<}{\sim}$}}~ 3 \times 10^{-5}$), the present day \he3 abundance found in standard models of galactic chemical evolution are not excessive and are consistent with the observed range of $1 - 5 \times 10^{-5}$ (Balser {\it et al.}~ 1994). However, even in these cases, there appears to be a problem with the abundance of \he3 as measured in meteorites, giving the pre-solar value of \he3. That is, on the basis of chemical evolution models, we expect more \he3 than is observed in the solar system (Olive {\it et al.}~ 1995; Galli {\it et al.}~ 1995; Tosi {\it et al.}~ 1995). In this paper, we will thus consider in turn, the possibilities that \he3 in the solar system has been depleted with respect to the galactic average; the accuracy of the measurement of solar \he3; and what we can learn from galactic chemical evolution. \section{Solar Depletion of \he3?} In principle, it is possible that the abundance of \he3 at the time of the formation of the solar system does not reflect the galactic average at that time. It is necessary to consider the degree to which element abundances in the solar system were affected by the explosions of supernovae in the solar neighborhood, immediately prior to the formation of the solar system (Reeves 1978; Olive \& Schramm 1982). This may be evidenced by several ``anomalous" isotopes (of Carbon, Oxygen and Neon) seen in cosmic rays. However, late contributions to the abundances of $^{16}$O and $^{20}$Ne, which are produced solely in Type II supernovae, may render the solar system isotopic ratios of these elements anomalous. That is, the solar abundances may not represent the true average galactic abundance. For the oxygen and neon isotopes, these differences may be as large as a factor of two. We now examine the possibility that the solar \he3 abundance may also not be representative of the true galactic abundance. {}From the observed anomalies in $^{26}$Al and $^{107}$Pd (Lee 1979; and references therein) and more recently $^{41}$Ca, and the short time scales associated with their half-lives ($\sim 10^6$ years), the element abundances in our solar system were probably affected by at least one supernova within that time period prior to formation. Even a single supernova explosion in a star forming region can have dramatic consequences on the element abundances of that region. As was argued by Olive \& Schramm (1982), a handful of the first few supernovae in an early OB association, is capable of producing nearly the entire observed solar abundance of $^{16}$O and $^{20}$Ne. Thus we would expect that the solar isotopic ratios such as $^{17,18}$O/$^{16}$O and $^{22}$Ne/$^{20}$Ne may be diluted with respect to the galactic average. It is therefore of interest to question whether or not the abundance of \he3 (which would be depleted in the ejecta of the first few supernovae in an association) is comparably depleted. To deplete \he3, we must require that a significant amount of material in the solar neighborhood underwent stellar processing prior to the formation of the solar system. Let us suppose that a fraction $f$ of the total initial gas of the association went into stars prior to the solar epoch. It is then reasonable to assume that a fraction $\sim 0.1f$ of the gas went into stars with masses greater than 10 $M_\odot$. (For example, with a Salpeter (1955) initial mass function (IMF) $\phi(m) \propto m^{-2.35}$ between, 0.1 and 100 $M_\odot$, the fraction is 12\%; for a Scalo (1986) mass function it may range from 5--15\% depending on the star formation rate (SFR)). If we denote by $X_*$ the mass fraction of some heavy element (such as O) ejected by massive stars, the total mass fraction of the element after the explosions of stars more massive than 10 $M_\odot$ which thus determines (and must be less than) the solar abundance, is then given by \begin{equation} X_f = {0.1fX_* + (1-f)X_i \over (1-f) + 0.1 f} < X_\odot \label{xf} \end{equation} where $X_i$ is the initial mass fraction of the element and we have assumed that $0.9f$ of the initial gas mass is still locked in stars. To maximize our estimate of \he3 destruction (as this will maximize our estimate for $f$), we can assume that $X_i = 0$. Solving for $f$, we have, \begin{equation} f < {X_\odot \over 0.1X_* + 0.9 X_\odot} \end{equation} For oxygen, $X_* \sim 0.1$ and $X_\odot \sim 0.01$, so that $f ~\mbox{\raisebox{-.7ex}{$\stackrel{<}{\sim}$}}~ 1/2$. Thus, we can cycle no more than 1/2 the mass of the association through stars prior to the formation of the solar system. Although a significant amount of gas may be cycled through stars, only a small fraction of \he3 depleted gas can be released back into the association. If we take $X$ to represent \he3 in Eq. (\ref{xf}), and now take $X_* = 0$ (which assumes that \he3 is totally destroyed in massive stars), and $X_i $ to be the primordial \he3 mass fraction, we find \begin{equation} X_f = {(1-f)X_i \over 1 - 0.9f} ~\mbox{\raisebox{-.7ex}{$\stackrel{<}{\sim}$}}~ .9 X_i \end{equation} This indicates that only about 10\% of the initial \he3 can be destroyed, even though changes in the heavy element abundances occur at a level of a factor of 2. It is possible, of course, to further deplete \he3 in the gas which forms the solar system at the expense of excessive metal production. Such overproduction of metals can perhaps be reconciled with \he3 depletion, if the heavy elements could somehow be expelled from the solar neighborhood. As Lattimer, Schramm \& Grossman (1977) pointed out, the bulk of the heavy element ejecta from supernovae can rapidly form into dust grains. These dust grains can behave like explosive ``shrapnel" and penetrate regions exterior to the association. This might allow the association itself and hence the solar system, to fail to show a large heavy element excess, even though the total heavy element enrichment would be part of the integrated galactic enrichment. This assumes, of course, that the entire association region is not totally disrupted by the supernovae explosion prior to the formation of the solar system. However, as was shown in Olive \& Schramm (1982), a significant amount of oxygen and neon is produced which should not be trapped in grains. Because of the behaviors of these elements, we believe that is unlikely that more than about 10\% of the \he3 present in the association could be destroyed before the formation of the solar system. An obvious recourse to resolving the problem of the overproduction of \he3 at the solar epoch, is to question the measurement of the solar \he3 abundance. In the next section, we examine the observational data on the solar abundances of D and \he3. \section{D and $^3$He in Pre-solar Nebulae} Because the crucial data in attempts to estimate the primordial D/H and $^3$He/H values come from solar system measurements, it is useful to examine critically the origin of these abundances and to attempt to provide an accurate estimate of their uncertainties. The determination of the solar abundances of D and \he3 involves $^3$He/$^4$He measurements both in meteorites and also directly in the solar wind. Direct D/H measurements are irrelevant for the Sun, since D is completely burned to $^3$He in the solar convective zone. Moreover, D/H measurements are difficult to interpret in planetary bodies (Earth, Jupiter, etc.); because D preferentially enters molecules relative to H, abundance determinations thus require a knowledge of complex chemical fractionation histories. However, because essentially all primordial D has been burned to $^3$He in the solar convective zone and because the convective zone is not hot enough to burn $^3$He, the solar wind measurement of $^3$He/$^4$He provides a measurement of the pre-solar abundance of ${{\rm D + ^3He} \over ^4 He}|_\odot$ by number. Solar wind measurements, made using foil collectors during Apollo lunar missions, yielded values for $^3$He/$^4$He ranging from 4 to 5.5 $\times 10^{-4}$. Geiss and Reeves (1972) and Bochsler \& Geiss (1989) (see also Geiss (1993) for a recent review) argue that the variation can be corrected for, and that the best solar wind ratio is ${\rm ^3 He/ ^4 He \vert _{sw} = 4.1 \pm 1 \times 10^{-4}}$ (where the error is statistical). This is in good agreement with the low temperature component emitted by carbonaceous chondrites in step-wise heating experiments (Black 1972; Weiler {\it et al.}~ 1991), for which ${\rm ^3 He/ ^4 He \vert _{sw} \simeq 4.5 \pm 1 \times 10^{-4}}$, and also with the ISEE-3 solar wind data (Coplan {\it et al.}~, 1984), which yields $4.4 \times 10^{-4}$. However, some fractionation in all the solar wind $^3$He/$^4$He measurements cannot be excluded, which would add an additional systematic error to the above value. The extreme value for $^3$He/$^4$He observed for the Apollo solar wind measurement of $5.5 \times 10^{-4}$ can not be excluded as a central value, hence a systematic uncertainty of $1.4 \times 10^{-4}$ for $^3$He/$^4$He cannot be excluded at present. The most recent measurement of $^3$He/$^4$He in the solar wind from the over-the-solar-pole measurements made with the SWICS instrument on the ULYSSES spacecraft gives \he3/\he4 $= (4.4 \pm 0.4) \times 10^{-4}$ (Bodmer {\it et al.}~ 1995). The pre-solar $^3$He/$^4{\rm He}\vert_\odot$ ratio is thought to be best measured in meteorites. Initially, Black (1971) proposed that the high temperature component emitted by step-wise heating experiments using carbonaceous chondrites (see also Eberhardt 1974) was the primordial component. $^3$He/$^4 {\rm He} \sim 1.5 \times 10^{-4}$. However, Weiler {\it et al.}~ (1991) have argued that this high temperature component is dominated by gas trapped in pre-solar grains (diamonds) which formed in locations far removed from the solar system. Weiler {\it et al.}~ propose that another gas component known as ``Q" is a better candidate for the primordial component. Fortunately, the difference in $^3$He/$^4$He between the high T carbonaceous chondrite component and Q is relatively small \begin{equation} ^3 He/^4 He \vert_Q = 1.6 \pm 0.04 \times 10^{-4} \end{equation} However, a potential interpretational (systematic) error persists, since neither Q nor diamonds nor the high T carbonaceous chondrite component has been unequivocally proven to represent $^3$He/$^4{\rm He}\vert_\odot$. Taking $^3$He/$^4{\rm He}\vert_q$ as $^3$He/$^4{\rm He}\vert_\odot$, but allowing for systematics to include the range of relevant meteoritic $^3$He/$^4$He values, yields $^3$He/$^4{\rm He}\vert_\odot = 1.6 \pm 0.04 \pm 0.3 \times 10^{-4}$. The pre-solar D is estimated by subtracting $^3$He/$^4{\rm He}\vert_\odot$ from the SWICS solar wind value. To convert to ratios relative to hydrogen requires multiplying by the number ratio of ${\rm ^4He/H \vert_\odot}$, which is estimated to be $0.09 \pm 0.01$ (note, this is 10\% lower than that used by Geiss 1993) from the best fit solar model $Y = 0.27$ (Turck-Chieze {\it et al.}~ 1988; Bahcall and Pinnsoneault 1992) with metallicity Z = 0.02. This yields \begin{equation} ({{D + ^3 He} \over H})_\odot = 4.1 \pm 0.6 \pm 1.4 \times 10^{-5} \end{equation} and \begin{equation} {{^3 He} \over H} \vert_{\odot} = 1.5 \pm 0.2 \pm 0.3 \times 10^{-5} \end{equation} and thus \begin{equation} {{D} \over H}_{\odot} = 2.6 \pm 0.6 \pm 1.4 \times 10^{-5} \end{equation} This latter number is in reasonable agreement with the ${\rm {HD} \over {H_2}} = 1 - 3 \times 10^{-5}$ ratio measured in Jupiter (Smith, Scherpp \& Barnes, 1989). Although planetary D ratios are subject to chemical fractionation, this is minimized for HD on Jupiter, since the bulk of the deuterium and hydrogen is in HD and H$_2$ there. However, molecular line blanketing does still allow for significant systematic errors. For this reason, Jupiter is still not the best source for a solar system D determination, but it does provide a consistency check. \section{Chemical Evolution} The solar system abundance of $^3$He is thus seen to be approximately a factor of two lower than predicted by even the more optimistic models of Galactic chemical evolution, which tend to yield abundance ratios at least as high as $3 \times 10^{-5}$ for \he3/H, when \he3 production in lower mass stars is included (Olive {\it et al.}~ 1995). In what follows, we will look at three different approaches to resolving the problem of excess solar \he3. We first consider possibilities for which the primordial value of D/H is low. A low initial D/H lowers \he3/H, as there is less D to be converted to \he3 in the pre-main-sequence evolution of stars. However, as we will show, one can not take arbitrarily low values of D/H (of course D/H is always bounded from below by the ISM measurements of D/H yielding D/H = $1.6 \pm 0.09 {}^{+0.05}_{-0.1}$ (Linsky {\it et al.}~ 1993, 1995)), since some amount of deuterium destruction necessarily accompanies the production of heavy elements in the galaxy. We then consider ``higher" values of D/H, which require some dramatic changes to simple models of chemical evolution, such as an increased production of massive stars in the early galaxy as well as metal enriched outflow. We will also examine some remaining alternatives regarding the stellar production of \he3. Note however, that there may be a quite disturbing dispersion of D/H in the local ISM which would complicate the analysis (Vidal-Madjar 1991; Ferlet 1992, Linsky private communication) As was noted earlier, the questions concerning high verses low D/H may become moot, if the determinations of primordial D/H in quasar absorption systems yield a single consistent value. To date there are three measurements of D/H in quasar absorption systems. Two (in the same system) yield a high value for D/H $\approx 1.9-2.5 \times 10^{-4}$ (Carswell {\it et al.}~ 1994; Songaila {\it et al.}~ 1994), while the third (in a different system yields a significantly lower value, D/H $\approx 1-2 \times 10^{-5}$. It is clear that, on the basis of these measurements, we can not with confidence claim any knowledge of the primordial abundance of deuterium. Indeed, it has been argued ( Levshakov \& Takahara 1995) that measurements of this type may not be able to determine D/H to better than an order of magnitude. In other words, they would expect a large dispersion in the observational data. Is this what we are seeing? Interestingly enough, the two values for D/H identified above are in some respects both beneficial and detrimental to big bang nucleosynthesis. The high value of D/H corresponds to a value for the baryon-to-photon ratio $\eta \simeq 1.5 \times 10^{-10}$ (Walker {\it et al.}~, 1991). Consequences of this high D/H were recently discussed in Vangioni-Flam \& Cass\'e (1995). With regard to the other light elements produced in big bang nucleosynthesis, the low value for $\eta$ corresponds to a \he4 mass fraction $Y_P \simeq 0.23$, which is in remarkable agreement with what one expects from the data on \he4 from extragalactic H\thinspace{$\scriptstyle{\rm II}$}~regions (Olive \& Steigman 1995; Olive \& Scully 1995). \li7/H is predicted to be around $2 \times 10^{-10}$ which is also compatible within errors, with recent data (Molaro {\it et al.}~ 1995). The problem occurs with the evolution of \he3, when \he3 production is included (note that models of chemical evolution can be constructed which can account for the necessary D/H destruction in this case). In Olive {\it et al.}~ (1995), it was found that the abundance of \he3 at the time of solar system formation could be high by as much as a factor 10. Even in the absence of \he3 production, it was found that massive stars were required to destroy at least 90\% of their initial D + \he3, in order to reproduce the solar and ISM values of \he3. This amount of destruction is excessive, even for the most massive stars (Dearborn, Schramm \& Steigman, 1986). On the other hand, the low value of D/H between 1 and 2 $\times 10^{-5}$ corresponds to a value of $\eta \approx 7 - 9 \times 10^{-10}$. In contrast to the high D/H case, we would expect a much milder problem with \he3 (to be discussed below). However, now the \he4 mass fraction is predicted to be $Y_P > 0.249$, a value larger than most of the \he4 measurements (Pagel {\it et al.}~, 1992; Skillman {\it et al.}~ 1995) in extragalactic H\thinspace{$\scriptstyle{\rm II}$}~ regions, which already contain some non-primordial \he4. (However, again, possible systematic errors can not be excluded Copi {\it et al.}~, 1995a; Sasselov \& Goldwirth, 1995.) In addition, \li7/H is expected to be $> 5 \times 10^{-10}$ requiring a significant amount of \li7 depletion, contrary to what one expects (Steigman {\it et al.}~ 1993) from the positive measurements of \li6 in halo stars (Smith Lambert \& Nissen, 1992; Hobbs \& Thorburn, 1994). Furthermore, as we will next show, a minimal amount of D destruction is demanded for consistency with the observed level of heavy element production in the Galaxy. A completely flat evolution for D is probably excluded on these grounds. The classical constraints on galactic evolution are characterized by varying degrees of stringency. Among these, the trends in [Fe/H] with time are easily satisfied, since the age-metallicity relation suffers from a large dispersion over the observed age range (Edvardsson {\it et al.}~ 1993; Nissen 1995). The [O/Fe] vs [Fe/H] relationship is mainly sensitive to the stellar yields and not to the different histories of star formation (assuming a constant IMF). The metallicity distribution of disk stars is far from being definitely established. Indeed, much work is needed before a clear picture of the metallicity distribution can be reached (e.g. Olsen 1994; Cayrel, private communication). Information on metallicities, ages and kinematics, with the same high accuracy as obtained by Edvarsson {\it et al.}~ (1993), is needed for a much larger stellar sample. Moreover, Grenon (1989, 1990) remarks that the radial migration of stars in the Galaxy can blur the local metallicity distribution. Other global characteristics which should be considered are the gas fraction, $\sigma$, the overall metallicity, Z, and individual abundance ratios (Fe/H, O/H,...) at solar birth and in the present ISM. To the list of constraints, we must also add the D/H and \he3/H ratios at solar birth and at present time, in relation to the primordial value. Indeed, since primordial nucleosynthesis is much more constrained than galactic evolution, it is reasonable to harmonize the second to the first, and not the contrary (as has sometimes been done recently). Many models have been proposed to follow the chemical evolution of the Milky Way, invoking, for example, a prompt initial enrichment ( Truran \& Cameron 1971), infall of primordial material (Timmes {\it et al.}~ 1995; Fields 1995), metal enriched infall originating from the halo (Ostriker and Thuan 1975), and early massive star formation ((Larson 1986; Wyse and Silk 1987). Studies of galactic chemical evolution remain in their infancy, however, since we do not yet have good theories of galaxy formation and star formation. It would be unwise, for the sake of simplicity, to limit the investigation to ``classical" models under the pretext that they have been widely used. In effect, if the high primordial D/H ratio is confirmed, special models leading to a strong D destruction avoiding overproduction of \he3 and Z will be required. An alternative way of looking at variations from the galactic mean has been carried out by Copi, Schramm \& Turner (1995b) looking at the stochastic variations from galactic evolution models. Their conclusions concerning the allowed range of primordial D and \he3 are similar to, and compatible with, those we discuss here. \subsection{Low D/H} We will first explore the possible consequences of a very low primordial value of D/H and examine the extent to which a low D/H could explain the apparent flatness of the \he3/H evolution in the Galaxy. We begin by estimating the minimum possible amount of D/H destruction. In simplified models of galactic chemical evolution, it is possible to derive some analytic relations between abundances, yields, the gas fraction, and the IMF, if one assumes the instantaneous recycling approximation (that is, that the enriched mass that is ultimately to be ejected from a star is incorporated into the ISM at the time of formation of the star, in contrast to its appropriate delayed entry at the end of the star's lifetime). Indeed, the degree to which deuterium is destroyed can be expressed simply by (Ostriker and Tinsley 1975) \begin{equation} {\rm \frac{D}{~D_p}} = \sigma^{R/(1-R)} \label{D} \end{equation} where $\sigma$ is the gas mass fraction and the return fraction, $R$, is given by \begin{equation} R = \int_{M_1}^{M_{sup}} (M-M_{rem}) \phi(M) dM \label{R} \end{equation} In (\ref{R}), $M_1$ is the main-sequence turnoff mass (normally a function of time), $M_{sup}$ is the upper mass limit for star formation, and $M_{rem}$ is the remnant mass. It is also possible to express the metallicity in terms of the gas mass fraction and the yields of metals in stars (Searle \& Sargent 1972) \begin{equation} Z = {P_Z \over (1 - R)} \ln \sigma^{-1} \label{Z} \end{equation} where \begin{equation} P_Z = \int_{M_1}^{M_{sup}} ({M_Z \over M}) M \phi(M) dM \end{equation} and $M_Z/M$ is the mass fraction ejected in metals. Equations (\ref{D}) and (\ref{Z}) can be combined yielding \begin{equation} {\rm {D \over ~D_p}} = e^{-{ZR \over P_Z}} \end{equation} As one can see from Eq. (\ref{D}), a low primordial value for D/H, will require a small return fraction, $R$. In principle, one can easily adjust the IMF to yield a small value for $R$. However, because of the similarity in the definitions of $R$ and $P_Z$, their ratio is almost independent of the details of the IMF. Thus ${\rm D/D_p}$ near unity, implies a metallicity much less than solar. The interdependence between deuterium and metallicity can be seen in Figure 1. In order to reach solar metallicity at the time the solar system formed, we require a deuterium destruction factor of at least 1.6, implying that D/H$_p ~\mbox{\raisebox{-.7ex}{$\stackrel{>}{\sim}$}}~ 2.5 \times 10^{-5}$. We note that this factor is somewhat dependent upon the assumed yields for the heavier elements. For example, this limit was obtained using the stellar yields of Woosley \& Weaver (1993), whereas had we used the yields of Maeder (1992), which allow for more heavy element production in the mass range from 9 - 11 M$_\odot$, the minimum destruction factor could be lowered to about 1.3. It is worth noting that, beyond the uncertainties of the yields which are essentially related to those associated with the $^{12}$C($\alpha,\gamma$)$^{16}$O reaction rate, the lower mass limit of the stellar progenitor of the Type II supernovae which synthesize the heavy elements is influential because of the preference in the IMF towards lower mass stars. Indeed the limit is greatly increased as the slope of the IMF is decreased. All of these effects can be seen in Figure 1, where we have plotted (for various choices of the parameters which govern the SFR) the metallicity at the solar epoch in units of solar metallicity, Z/Z$_\odot$, as a function of the ratio of the {\em present} deuterium abundance to the primordial one, thus indicating the total deuterium destruction factor for a variety of galactic evolution models. We have chosen a SFR proportional to the mass in gas, and an IMF, $\phi(m) \propto m^{-2.7}$; shown here by the upper set of points denoted by circles (yields from Maeder (1992)) and crosses (yields from Woosley \& Weaver (1993)). For the lower set of points, a steeper IMF, $\phi(m) \propto m^{-3}$ was chosen. In each case, the lower mass limit of the IMF was 0.4 M$_\odot$ (lowering this choice to 0.1 M$_\odot$ would further lower the curves). It is important to note that, when \he3 production is taken into account, even the modest deuterium destruction factor (of 1.6), yields an overproduction by about a factor of 2 in solar \he3. \subsection{Higher D/H} In this section, we will consider an alternative to low primordial D/H and rely on more distinctive models of galactic chemical evolution to resolve the problem concerning the solar \he3 abundance. As we have stated earlier, the choice of a higher value for primordial D/H alleviates some of the pressure in matching the BBN calculations to the observational determinations of \he4 and \li7. Clearly, the higher the value we choose for primordial D/H, the more difficult it will be to keep \he3 under control. We choose specifically the value D/H$|_{\rm p} = 7.5 \times 10^{-5}$ which corresponds roughly to the \li7 trough and is in modest agreement with \he4 (at the 2 $\sigma$ plus systematics level). The models we consider below specifically involve mass outflow. Open galactic models have been considered in the past (Tinsley 1980, Tosi 1988), but infall has been invoked more often than outflow. Formally, the two reverse processes are included in the general formalism of chemical galactic evolution (Tinsley 1980) and cosmochronology (Cowan {\it et al.}~ 1989). There is clear evidence for galactic winds in external galaxies, even for spirals (Wang {\it et al.}~ 1995), and particularly those experiencing bursts of star formation, whereas evidence for significant infall of extragalactic matter are meager (Murphy {\it et al.}~ 1995). Of course this reflects only constraints arising from the present state of the solar vicinity, and proves nothing about the early galaxy. Outflow has its own merit: as we will see below, it will help to explain very high destruction factors of D (Vangioni-Flam \& Cass\'{e} 1995) while, if necessary, avoiding, an overproduction of metals. At the same time it could reduce the rise of the \he3/H ratio. De Young \& Heckman (1994) proposed that energy from supernova explosions and stellar driven winds results in blowing portions of the ISM containing enriched material out of the galaxy. Different kinds of outflows can be imagined, which vary considerably in their durations, intensities and compositions. For our present purposes, it is sufficient to distinguish whether the outflowing matter consists solely of the ejecta of massive stars, or rather whether it is composed of normal ISM material being blown out by supernova driven winds. In Cass\'{e} {\it et al.}~ (1995), we will return to these distinctions in greater detail. The three specific ingredients that must be added to canonical galactic evolutionary models in order to obtain significant D destruction without the overproduction of \he3 are: (i) an early phase of massive star formation, which presents the advantage of destroying D and \he3 rapidly; (ii) a galactic wind related to the corresponding SNII rate, which limits the rise of Z and \he3, leading to an even more pronounced decrease of D; and (iii) possible modifications of stellar models, leading to an efficient destruction of \he3, especially in low mass stars. One way we have found in which the solar value of \he3 may be lowered is to assume that the IMF prior to the formation of the solar system was skewed more toward massive star formation. The presence of fewer lower mass stars reduces \he3 production, while more massive stars ultimately return only a fraction of the \he3 present during the pre-main sequence phase to the ISM. We therefore consider models which begin with an IMF favoring more massive stars early on galactic history but resemble a more normal IMF at later times. The problem that we immediately encounter is that the emphasis on more massive stars results in an overproduction of heavy elements, such as $^{16}$O. We have found that the (closed box) models which are most successfully in keeping \he3 flat, while destroying enough D, also overproduce $^{16}$O by a factor of $\sim$10. This problem is alleviated by including outflow. Indeed, McCray \& Snow (1979) have shown that supernovae can generate ``chimneys," which can directly transport much of the heavy element rich supernova debris out of the galaxy. We have therefore included ``enriched'' (relative to the ISM) outflow in our models, both to help solve the heavy element overproduction problem and to obtain a flatter \he3/H evolution. In order to simulate this effect, we have incorporated outflow into our models at a rate proportional to the rate of ejection of materials from supernovae. Since massive stars can lose large amounts of \he3 depleted outer material via winds before they explode, it is certainly possible for them to deplete \he3 in their surrounding ISM material and eject their metals out of the Galaxy. We allow the outer (hydrogen) envelope of the star which is deficient in \he3 to return via winds to the ISM. Then, in order to maximize the possible effect of an outflow which is tied to the ejecta of exploding stars (M $>$ 8-10 M$_\odot$), a fraction of the core is then expelled from the Galaxy. Such models provide a natural way to understand the heavy element abundances in the X-ray gas observed in clusters of galaxies (typically, the heavy element enriched outflow is produced by elliptic galaxies, (see e.g. Elbaz, Arnaud, Vangioni-Flam 1995)). It might also be noted that early expulsion of metal rich supernova ejecta is even easier in merger models, where the early galactic building blocks have a lower mass. Due to the epoch of more massive star formation at earlier times, D is generally very efficiently destroyed in these models. We should note at this point that our assumptions concerning winds from massive stars prior to the supernova stage may be inappropriate at early galactic epochs. The rate of mass loss is generally expected to be dependent upon the initial metallicity of the star (Maeder, Lequeux, \& Azzopardi 1980; Maeder \& Meynet 1994). Expectations from theoretical studies are generally consistent with e.g. trends in the frequency of Wolf-Rayet stars, as inferred from studies of the Magellanic Clouds (Massey {\it et al.}~ 1995). This suggests that the fraction of the \he3 depleted outer envelopes of massive stars that is returned to the ISM via winds (prior to supernova-triggered mass ejection) in low metallicity populations may be significantly reduced. We stress however, that our aim here is to see how efficiently the evolution of the \he3 abundance can be held relatively flat over the history of the Galaxy. As we will see, we find only modest success despite rather poignant assumptions. An obvious observational constraint on our choice of an IMF that is skewed toward more massive star production at early times, $\phi(m) \propto m^{-(1.25 + O/O_\odot)}$, is provided by its consistency with the present day IMF that results from our model. Fig. 2 shows a comparison of the observed and modeled present day IMF. The observed values are taken from Scalo (1986) and are in good agreement with our model for the more massive stars. This is as expected, since the more numerous massive stars formed early on have long since died out. It is be useful to compare the results we have obtained here with those of our previous study, in which we considered a more standard model. In model 1 of Olive {\it et al.}~ (1995), we chose a SFR, $\psi = 0.25\sigma$, with an IMF, $\phi(m) \propto m^{-2.7}$ between 0.4 and 100 M$_\odot$. The \he3 abundance at the solar formation epoch (taken to be at $t = 9.4$ Gyr) was \he3/H = 5.7 $\times 10^{-5}$, rising to 8.8 $\times 10^{-5}$ today. Infall was not included. For a primordial ratio D/H = $7.5 \times 10^{-5}$, the results for D/H and \he3/H as a function of time, for a model with an IMF which favors massive stars early, and contains enriched outflow in which the rate is proportional to the ejection rate, is compared with model 1 of Olive {\it et al.}~ (1995) in Figure 3. The outflowing gas contains only material below the outer envelope, while the latter \he3 depleted material is returned to the ISM. In this model, the SFR is $\psi = 0.26 M_{\rm gas}$, and the fraction of outflowing gas is 90\% of the supernova ejecta. The IMF is now extended down to 0.1 M$_\odot$, to help keep the metallicity and gas mass fraction reasonably low. We view this as a rather extreme model, in which a considerable amount of enriched material has been expelled from the galaxy. Indeed, we impose a limit, arising from the observed metallicity of hot X-ray gas in clusters, on the amount of metals expelled by outflow to be less than 20 times the amount of metals in the galaxy. This imposes a constraint on the fraction of outflowing gas (90\% in this case). As one can see from Fig. 3, our present model, which is based on an IMF skewed towards massive stars early on and contains enriched outflow, reduces the abundance of \he3 by a factor of about 2, relative to the standard case with a normal IMF and no outflow. Parameters of the model have been chosen such that the degree of deuterium destruction is comparable (and agrees with the data) in the two cases. However, although the present \he3 abundance is acceptable \he3/H$|_o = 5.1 \times 10^{-5}$, the solar abundance is still high by a factor of slightly over two; \he3/H$|_\odot = 3.7 \times 10^{-5}$. While this represents a definite improvement, it can not be regarded as a solution to the problem. Although it appears from Figure 3, better agreement with the solar data is possible if one assumes a lower time for the formation of the solar system, the model must be adjusted to destroy D on a faster time scale. For example, with $\psi = .34M_{\rm gas}$ the evolution of deuterium matches the solar (and present-day) observations at $t = 6$ Gyr (corresponding to an age of the Galaxy of 10.6 Gyr), but now \he3/H$|_\odot$ $\simeq 3.1 \times 10^{-5}$ and the present abundance is 4.9 $\times 10^{-5}$; a further improvement, but solar \he3 is still too high. Of course as is well known, the problem concerning \he3 is also alleviated somewhat by going to higher values of $\eta$. In model 3 of Olive {\it et al.}~ (1995), we assumed a primordial abundance of deuterium of D/H = 3.5 $ \times 10^{-5}$. In this case \he3/H$|_\odot = 3.4 \times 10^{-5}$, an overproduction by a factor greater than two. The present \he3 was also slightly high, \he3/H = $6 \times 10^{-5}$. In models with outflow as described above, these numbers are reduced to \he3/H$|_\odot = 2.3 \times 10^{-5}$ and \he3/H = $3.2 \times 10^{-5}$ today. Before, we move on, we wish to stress that the problems concerning \he3 that we are discussing here, prevail only because we are including the production of \he3 in low mass stars. When such production is ignored there is no problem in matching the solar and ISM data for both D and \he3 in models of these types as was shown by Vangioni-Flam, Olive \& Prantzos (1994). The crises in big bang nucleosynthesis claimed by Hata {\it et al.}~ (1995) is only a crises because of the limit on the degree of \he3 destruction they allowed. Although the final \he3 abundance in a given star relative to the initial D + \he3 abundance, usually called $g_3$, is always larger than 0.25 as assumed by Hata {\it et al.}~, even simple models such as the type considered here (without outflow) and in Vangioni-Flam, Olive \& Prantzos (1994) have an effective $g_3$ which is lower than 0.25 vitiating the purported crises. \subsection{Alternatives} A critical consideration with regard to the establishment of any realistic constraints on cosmological D and D + $^3$He is that associated with $^3$He production in low mass stars. Essentially all early estimates of D and $^3$He constraints on cosmology (see, e.g., Truran \& Cameron 1971; Rood, Steigman, \& Tinsley 1976) were based upon the stellar evolution models of Iben (1967 a,b), for which analytical fits to the detailed model characteristics were subsequently provided by Iben \& Truran (1978). The problem of $^3$He then is simply the fact that, with the use of the Iben \& Truran (1978) prescriptions, $^3$He production in stars in the mass range $\sim$ 1-3 M$_\odot$ is sufficient to overproduce $^3$He in Galactic chemical evolution models (Olive {\it {\it et al.}~} 1995; Galli {\it et al.}~ 1995; Timmes \& Truran 1995), relative both to the solar system value of $^3$He and to the $^3$He concentration in the ISM at the present time (Balser {\it {\it et al.}~} 1994). Further strong confirmation of this behavior has been provided by recent stellar evolution calculations (Vassiliadis \& Wood 1994; Weiss {\it et al.}~ 1995). It would seem to be necessary either to utilize rather extreme assumptions regarding the history of our Galaxy or to identify some significant problem in stellar evolution theory. An interesting recent paper by Wasserburg, Boothroyd, \& Sackmann (1995) has called attention to the fact that the long-standing problems associated with understanding both low $^{12}$C/$^{13}$C ratios in low mass red giant branch stars and low $^{18}$O/$^{16}$O ratios in asymptotic giant branch stars can be resolved, with the assumption of the occurrence of deep circulation currents extending below the bottom of the standard convective envelope. A concomitant of this process of ``cool bottom burning'' is the destruction of $^3$He. In particular, for the case of a 1 M$_\odot$ star, their models predict that after a 1$^{st}$ dredge-up $^3$He enhancement of a factor $\sim$ 6, cool bottom processing acts to reduce the $^3$He concentration by a factor $\sim$ 10, yielding a net \underbar{depletion} of $^3$He by a factor $\sim$ 2. If this model is indeed correct, this would aid substantially in the problem of $^3$He overproduction, with which we are so concerned in this paper. To test the effect of such a reduction in the \he3 yields, we incorporated the results of Wasserburg {\it et al.}~ (1995) by lowering the Iben \& Truran (1978) yields of \he3 at 1 M$_\odot$ by a factor of 10. For an initial deuterium abundance of 7.5 $\times 10^{-5}$, this corresponds to a $g_3 = 0.27$. We reduced the degree to which the Iben \& Truran yields were modified at higher masses such that, at M $>$ 3 M$_\odot$, we once again were using the Iben \& Truran yields. The results of such a reduction in model 1 of Olive {\it et al.}~ (1995) are shown in Figure 4. Here the evolution of D/H and \he3/H are shown in model 1 with both the Iben \& Truran yields and the reduced yields. Even in this case, there remains a mild overproduction of \he3 by a factor of about 2. That is, at $t = 9.4$ Gyr, \he3/H = 3 $\times 10^{-5}$. In Figure 5, we show the effect of the reduced \he3 yields in the model with outflow discussed in the previous section. Here finally, we find a value for \he3/H at the solar epoch which is perhaps acceptable. At $t = 9.4$ Gyr, \he3/H = 2.4 $\times 10^{-5}$. A further improvement is possible by considering models which evolve on shorter time scales as discussed above. An obvious problem with the reduction in \he3 yields at low stellar masses is the observation of high $^3$He concentrations in planetary nebula ejecta (Rood {\it {\it et al.}~} 1992, 1995), which would seem to confirm the predictions of the more standard models for the evolution of low mass stars along the giant branch. It is clear that this issue must be resolved before a more definitive statement can be made with respect to Galactic evolution constraints on the primordial abundances of D and $^3$He. A further question of interest is that concerning the composition of the matter involved in ``mass infall,'' during the later stages of evolution of our Galaxy. In this context, we note that while we have not considered such infall models in this paper (see, e.g., Olive {\it et al.}~ 1995), they may provide plausible alternative solutions to the \he3 problem. The implications of infall of matter of primordial composition of the light elements D. $^3$He, and $^7$Li are certainly quite different from those of processed matter, which may generally be expected to be metal enriched and deuterium depleted. Infall of primordial material is generally beneficial, with the adoption of the lower primordial D abundance, while infall of D- and $^3$He-depleted matter improves the situation for the case of a higher primordial D abundance. The fact that the nature and origin of such infalling material is presently uncertain, makes it necessary to treat its composition as an additional parameter. This problem is further complicated by the fact that it is even possible for the in falling gas both to be metal enriched and to have an essentially primordial composition of D and $^3$He. Such could occur if, for example, the ejecta of the first generation of massive stars in the halo of our Galaxy were lost to the surrounding intergalactic medium. The ejecta of stars of M $>$ 10M$_\odot$ collectively represents $\sim$ 10 \% of the initial mass formed into stars (e.g., for a Salpeter IMF over the range 0.1-100 M$\odot$) and is characterized by a metal abundance $>$ 10 Z$_\odot$. Assuming $\sim$ 10$^{10}$ M$_\odot$ of early halo star formation would yield $\sim$ 10$^9$ M$_\odot$ of metal enriched gas ejected, which could contaminate $\sim$ 10$^{10}$ M$_\odot$ to solar metallicity and yet have deuterium at a level of only $\sim$ 0.9 its primordial value. \section{Conclusions} What can we conclude from this analysis? We have shown that, unlike the the abundances of some of the heavier elements such as oxygen and neon which can differ by as much as a factor of two locally relative to their average galactic abundance by prior supernova in the solar neighborhood, the local \he3 abundance could only have been affected by at most 10\%. It also appears that the \he3 data from a variety of sources is consistent and yields a value \he3/H$|_\odot = 1.5 \times 10^{-5}$ for the presolar \he3 abundance. Standard models of galactic chemical evolution yield an excess of \he3 at the solar epoch by a factor which ranges from 2 to 12 depending on the assumed primordial value for D/H. For models with D/H = $7.5 \times 10^{-5}$ initially, the factor nearly 4 excess in \he3 can be brought down to an excess of about 2 in models which favor massive stars early on, and include the possibility for a substantial amount of metal enriched outflow. In such models, the solar \he3 abundance is brought down to nearly acceptable levels when primordial D/H $< 3.5 \times 10^{-5}$. Finally, we considered the possibility that part of the problem may lie in the stellar yields of \he3. Though it appears that the cut in yields suggested by Wasserburg {\it et al.}~ (1995) may not in itself be sufficient to lower the solar \he3 abundance, that reduction in conjunction with chemical evolution models may. We feel justified in claiming that any apparent ``crises" in big bang nucleosynthesis is rather a (potential) problem for chemical evolution and/or stellar evolution. \bigskip \noindent {\bf Acknowledgments} We would like to thank J. Audouze, C. Copi, R. Lewis, R. Pepin, G. Steigman F. Timmes, and M. Turner for helpful conversations. The work of KAO was supported in part by DOE grant DE-FG02-94ER-40823. The work of DNS was supported in part by the DOE (at Chicago and Fermilab) and by the NASA through NAGW-2381 (at Fermilab) and a GSRP fellowship at Chicago. The research of JWT has been supported by the National Science Foundation under grant NSF AST 92-17969 and by the National Aeronautics and Space Administration under grant NASA NAG 5-2081. The work of EV-F was supported in part by PICS $n^o$114, ``Origin and Evolution of the Light Elements", CNRS. \newpage \begingroup \section*{\large \bf References \bibitem Bahcall, J.N. \& Pinsonneault, M.H. 1992, Rev. Mod. Phys., 64,885 \bibitem Balser, D.S., Bania, T.M., Brockway, C.J., Rood, R.T., \& Wilson, T.L. 1994, ApJ, 430, 667 \bibitem Black, D.C. 1971, Nature Phys. Sci., 234, 148 \bibitem Black, D.C. 1972, Geochim. Cosmochim. Acta, 36, 347 \bibitem Bochsler, P. \& Geiss, J. 1989, in {\it Proc. Yosemite Conf. on Outstanding Problems in the Solar System} p.133 \bibitem Bodmer, R., Bochsler, P., Geiss, J., Von Steiger, R., \& Gloeckler, G. 1995, Spa. Sci Rev., 72, 61 \bibitem Carswell, R.F., Rauch, M., Weymann, R.J., Cooke, A.J. \& Webb, J.K. 1994, MNRAS, 268, L1 \bibitem Cass\'{e}, M., Olive, K.A., Scully, S.T., \& Vangioni-Flam, E. 1995, in preparation \bibitem Charbonnel,C. 1994, A \& A, 282, 811 \bibitem Copi, C. Schramm,D.N. \& Turner, M.S. 1995a, Science, 267,192 \bibitem Copi, C. Schramm,D.N. \& Turner, M.S. 1995b, {ApJ Lett} submitted \bibitem Coplan, M.A., Ogilvie, K.W., Bochsler, P., \& Geiss, J. 1984, Solar Phys., 93, 415 \bibitem Cowan, J.J., Thielemann, F.-K., \& Truran, J.W. 1987, ApJ, 323, 543 \bibitem Dearborn, D. S. P., Schramm, D., \& Steigman, G. 1986, ApJ, 302, 35 \bibitem Delbourgo-Salvador, P., Gry, C., Malinie, G., \& Audouze, J. 1985, A\&A, 150, 53 \bibitem De Young, D.S. \& Heckman, T.M. 1994, ApJ, 431, 598 \bibitem Eberhardt, P. 1974, Earth Planet. Sci., 24, 182 \bibitem Edvardsson, B., Anderson, J., Gustafson, B., Lambert, D.L., Nissen, P.E., \& Tomkin, J. 1993, A \& A, 275, 101 \bibitem Elbaz, D., Arnaud, M., \& Vangioni-Flam, E. 1995, A \& A, in press \bibitem Fields, B. 1995, ApJ, in press \bibitem Galli, D., Palla, F. Ferrini, F., \& Penco,U. 1995, ApJ, 433, 536 \bibitem Geiss, J. 1993, in {\it Origin and Evolution of the Elements} eds. N. Prantzos, E. Vangioni-Flam, and M. Cass\'{e} (Cambridge:Cambridge University Press), p. 89 \bibitem Geiss, J. \& Reeves, H. 1972, A \& A, 18,126 \bibitem Grenon, M. 1989, Astrophys. \& Science, 156, 29 \bibitem Grenon, M. 1990, in {\it Astrophysical Ages and Dating Methods}, ed. E. Vangioni-Flam {\it et al.}~ (Ed. Frontieres, Paris), p. 153 \bibitem Gry, C., Malinie, G., Audouze, J., \& Vidal-Madjar, A. 1984, in Formation and Evolution of Galaxies and Large Scale Structure in the Universe, eds. J. Audouze \& J. Tran Tranh Van (Reidel, Dordrecht) p 279 \bibitem Hata, N., Scherrer, R.J., Steigman, G., Thomas, D., Walker, T.P., Bludman, S., \& Langacker, P. 1995, preprint hep-ph/9505319 \bibitem Hobbs, L. \& Thorburn, J. 1994, ApJ, 428, L25 \bibitem Hogan, C.J. 1995, ApJ, 441, L17 \bibitem Iben, I. 1967a, ApJ, 147, 624 \bibitem Iben, I. 1967b, ApJ, 147, 650 \bibitem Iben, I. \& Truran, J.W. 1978, ApJ, 220,980 \bibitem Larson, R.B. 1986, MNRAS, 218, 409 \bibitem Lattimer, J., Schramm, D.N., \& Grossman, L. 1977, ApJ, 214, 819 \bibitem Lee, T. 1979, Rev. Geophys. Space Phys., 17, 1591 \bibitem Levshakov, S.A., \& Takahara, F. 1995, preprint \bibitem Linsky, J.L., Brown, A., Gayley, K., Diplas, A., Savage, B. D., Ayres, T. R., Landsman, W., Shore, S. N., Heap, S. R. 1993, ApJ, 402, 694 \bibitem Linsky, J.L., Diplas, A., Wood, B.E., Brown, A., Ayres, T. R., Savage, B. D., 1995, ApJ, in press \bibitem Maeder, A. 1992, A \& A 264, 105 \bibitem Maeder, A. \& Meynet, G. 1994, A\&A, 287, 803 \bibitem Maeder, A., Lequeux, J., \& Azzopardi, M. 1980, A\&A, 90, L17 \bibitem Massey, P., Lang, C.C., DeGiola-Eastwood, K., \& Garmany, C.D. 1995, ApJ, 438, 188 \bibitem McCray, R. \& Snow, T.P. 1979, ARAA, 17, 213 \bibitem Molaro, P., Primas, F., \& Bonifacio, P. 1995, A \& A, 295, L47 \bibitem Murphy, E.M., Lockman, F.J. \& Savage, B.D. 1995, ApJ, 447, 642 \bibitem Nissen, P.E. 1995, in IAU Symposium 164, {\it Stellar Populations}, to be published \bibitem Olive, K.A., Rood, R.T., Schramm, D.N., Truran, J.W., \& Vangioni-Flam, E. 1995, ApJ, 444, 680 \bibitem Olive, K.A., \& Schramm, D.N. 1981, ApJ, 257, 276 \bibitem Olive, K.A., \& Scully, S.T. 1995, Int. J. Mod. Phys. A, in press \bibitem Olive, K.A., \& Steigman, G. 1995, ApJ S, 97, 49 \bibitem Olsen, E.H. 1994, A\&A Supp, 104, 429 \bibitem Ostriker, J.P., \& Thuan, T.X. 1975, ApJ, 202, 353 \bibitem Ostriker, J.P., \& Tinsley, B. 1975, ApJ, 201, L51 \bibitem Pagel, B E.J., Simonson, E.A., Terlevich, R.J. \& Edmunds, M. 1992, MNRAS, 255, 325 \bibitem Reeves, H. 1978, in {\it Protostars and Planets}, ed. T. Gehrels (Tucson:University of Arizona Press) \bibitem Rood, R.T., Bania, T.M., \& Wilson, T.L. 1992, Nature, 355, 618 \bibitem Rood, R.T., Bania, T.M., Wilson, T.L., \& Bania, D.S. 1995, in {\it the Light Element Abundances, Proceedings of the ESO/EIPC Workshop}, ed. P. Crane, (Berlin:Springer), p. 201 \bibitem Rood, R.T., Steigman, G. \& Tinsley, B.M. 1976, ApJ, 207, L57 \bibitem Salpeter, E.E. 1955, ApJ, 121, 161 \bibitem Sasselov, D. \& Goldwirth, D.S. 1995, ApJ, 444, L5 \bibitem Scalo, J. 1986, Fund. Cosm. Phys. 11, 1 \bibitem Searle, L. \& Sargent, W.L. 1972, ApJ, 173, 25 \bibitem Skillman, E., {\it et al.}~\ 1995, ApJ Lett (in preparation) \bibitem Smith, V.V., Lambert, D.L., \& Nissen, P.E., 1992, ApJ 408, 262 \bibitem Smith, Scherpp \& Barnes 1989, ApJ, 336, 167 \bibitem Songaila, A., Cowie, L.L., Hogan, C. \& Rugers, M. 1994 Nature, 368, 599 \bibitem Steigman, G., Fields, B. D., Olive, K. A., Schramm, D. N., \& Walker, T. P., 1993, ApJ 415, L35 \bibitem Timmes, F.X., \& Truran, J.W. 1995, preprint \bibitem Timmes, F.X., Woosley, S.E., \& Weaver, T.A. 1995, ApJSupp, 98, 617 \bibitem Tinsley, B.M. 1980, Fund. Cosmic Phys., 5, 287 \bibitem Tosi, M. 1988, A\&A, 197, 33 \bibitem Tosi, M., Steigman, G. \& Dearborn, D.S.P. 1995, in {\it the Light Element Abundances, Proceedings of the ESO/EIPC Workshop}, ed. P. Crane, (Berlin:Springer), p.228 \bibitem Truran, J.W., \& Cameron, A.G.W. 1971, ApSpSci, 14, 179 \bibitem Turck-Chi\`eze, S., Cahen, S., Cass\'e, M., \& Doom, C. 1988, ApJ, 335, 415 \bibitem Tytler, D. \& Fan, X.-M. 1995, BAAS, 26, 1424 \bibitem Vangioni-Flam, E., \& Audouze, J. 1988, A\&A, 193, 81 \bibitem Vangioni-Flam, E. \& Cass\'{e}, M. 1995, ApJ, 441, 471 \bibitem Vangioni-Flam, E., Olive, K.A., \& Prantzos, N. 1994, ApJ, 427, 618 \bibitem Vassiladis, E. \& Wood, P.R. 1993, ApJ, 413, 641 \bibitem Vidal-Madjar, A. 1991, Adv. Space Res., 11, 97 \bibitem Walker, T. P., Steigman, G., Schramm, D. N., Olive, K. A., \& Kang, H. 1991 ApJ, 376, 51 \bibitem Wang, Q.D., Walterbos, R.A.M., Steakley, M.F., Norman, C.A., \& Braun, R. 1995, ApJ, 439, 176 \bibitem Wasserburg, G.J., Boothroyd, A.I., \& Sackmann, I.-J. 1995, ApJ, 447, L37 \bibitem Weiler, R., Anders, E., Bauer, H., Lewis, R. \& Signer, P. 1991, Geochim \& Cosmochim Acta, 55, 1709 \bibitem Weiss, A., Wagenhuber, J., and Denissenkov, P. 1995, preprint \bibitem Woosley, S.E. \& Weaver, T.A. 1993, in {\it Supernova, Les Houches Summer School Proceedings, Vol. 54}, ed. S. Bludman, R. Mochkovitch, \& J. Zinn-Justin (Geneva: Elsevier Science Publishers), p. 100 \bibitem Wyse, R., \& Silk, J. 1987, ApJ, 313, L11 \par \endgroup \newpage \noindent{\bf{Figure Captions}} \vskip.3truein \begin{itemize} \item[] \begin{enumerate} \item[] \begin{enumerate} \item[{\bf Figure 1:}] The dependence of the metallicity produced as a function of the deuterium destruction factor D/D$_p$, for a large sample of models. The metallicity is plotted in solar units. The SFR used was $\psi \propto M_{\rm gas}$ where the constant of proportionality ranges from 0.01 to 1.0 The circles correspond to the choice of stellar yields from Maeder (1992) while the crosses correspond to the yields of Woosley \& Weaver (1993). A power law IMF was chosen with a slope of -2.7 for the upper two sets of points and -3.0 for the lower two sets. \item[{\bf Figure 2:}] The present day mass function of our adopted model as compared with the data from Scalo (1986). \item[{\bf Figure 3:}] The evolution of D/H and \he3/H as a function of time, for a standard model of galactic chemical evolution (solid line) and for one which favors massive stars early and includes metal enriched outflow (dashed line). Also shown are the values of these ratios at the time of formation of the sun, $t \approx 9.4$ Gyr, and today, for D/H (open squares) and \he3/H (filled circles). The present day \he3 abundance simply shows the range of observed values; the data point does not represent an average. The models were chosen so that D/H is destroyed by a total factor of 5, to the present. \item[{\bf Figure 4:}] As in Figure 3, for a standard model (solid) and for one in which the stellar yields of \he3 at low masses have been reduced (dotted). Deuterium is the same in both cases. \item[{\bf Figure 5:}] As in Figure 3, for the model with enriched outflow from Figure 3 (dashed) and for one with outflow in which the stellar yields of \he3 at low masses have been reduced (dotted). \end{enumerate} \end{enumerate} \end{itemize} \end{document}
\section{Introduction} In the study of the two-dimensional integrable models of quantum field theories and statistical models, the Yang-Baxter equation (YBE) plays essential roles in establishing the integrability and solving the models. In the field theories, the YBE provides a consistency condition for the two-body scattering amplitudes ($S$-matrices) in the multi-particle scattering processes since the scattering is factorizable. With unitarity and crossing symmetry, the YBE can determine the $S$-matrix completely, although not uniquely due to the CDD factor which is any function of the rapidity satisfying the unitarity and crossing symmetry conditions. This CDD factor is usually neglected under the the minimality assumption. In the statistical models, if the Boltzmann weights satisfy the YBE, row-to-row transfer matrices with different values of the spectral parameter commute each other so that the models are integrable. Recently there has been a lot of efforts in extending these approaches to models with boundaries. The main motivation is that these models can be applied to 3D spherically symmetric physical systems where $s$-wave element becomes dominant. One-channel Kondo problem, monopole-catalyzed proton decay are frequently cited examples. Also one can generalize the conventional periodic boundary condition of the statistical models to other types like the fixed and free conditions. The existence of the boundary adds new quantities like boundary scattering amplitudes and boundary Boltzmann weights, and one needs to extend the YBE to include these objects. The boundary Yang-Baxter equation (BYBE) (also known as the reflection equation) \cite{cher} plays the role of the YBE for the integrable statistical models \cite{skly,oth} and quantum field theories \cite{zam} in the presence of a boundary; it is the necessary condition for the integrability of these models. The equation takes the form \begin{equation} R_{1}(u)S_{12}(u^{'}+u)R_2(u^{'})S_{12}(u^{'}-u)=S_{12}(u^{'}-u)S_2(u^{'}) S_{12}(u^{'}+u)R_{1}(u) \end{equation} where $R_{1(2)}$ is the boundary scattering (or reflection) matrix in the auxiliary space $1(2)$ and $S_{12}$ is the solution to the YBE. In general, $R(u)$ need not be a ${\bf C}$-number matrix, so the equation may be taken as the defining relation for the associative algebra generated by the symbols $R(u)$ \cite{ryu}. This algebra possesses a very rich structure and has been found to be connected with braid groups \cite{kul1}, lattice current algebra \cite{alek}, twisted Yangian \cite{naz} and so on. Taking the quantum space of $R(u)$ to be trivial, the BYBE is a quadratic matrix equation which allows the matrix $R(u)$ to be solved for given $S(u)$. To date, several solutions of the BYBE have appeared in the form of vertex representation, far less is known however for the solution in the solid on solid (SOS) or restricted solid on solid (RSOS) representation \cite{kul,pea}. Also less clear is the vertex-SOS correspondence associated with this algebra. In particular there is no reason to expect the vertex-SOS transformation for the YBE continue to hold for this algebra. Therefore finding solution in the RSOS form may help to clarify the issue of the vertex-SOS correspondence. Moreover, the RSOS solution will reveal the special mathematical structure associated with this algebra when the deformation parameter is a root of unity. {}From a physical point of view, the solutions have applications in statistical mechanics and field theory. In the context of statistical mechanics, the solutions give rise to integrable SOS/RSOS models with boundaries where the ${\bf C}$-number solution of the BYBE provides the Boltzmann weights of the statistical models at the boundary. The first nontrivial case gives the tri-critical Ising model. The study of integrable statistical model with boundary will shed light on the issue of the dependence of the Casimir energy on the boundary and surface properties \cite{yung,bau}. {}From the field theory point of view, the solutions are relevant to the study of the restricted sine-Gordon model \cite{abl} and the perturbed (coset \cite{bl,an}) conformal field theory (CFT) \cite{zam1} with boundary. In this case, solutions to the BYBE are scattering matrices of the particles with the boundary. In this paper we construct the BYBE in the RSOS/SOS representation. The equation is studied for the diagonal and non-diagonal cases, and the most general trigonometric solutions are found up to an overall factor. This factor is then fixed using the boundary crossing and unitarity conditions \cite{zam} (up to the usual CDD ambiguity). We then construct an integrable RSOS/SOS model using the diagonal solution and propose an algebraic Bethe ansatz method to diagonalize the transfer matrix of the SOS model. \section{Solutions to the boundary Yang-Baxter equation} \subsection{Generalities} In this section we solve the BYBE for the $\mbox{RSOS}(p)\;;$ $p=3,4\ldots$ scattering theory. The $\mbox{RSOS}(p)$ scattering theory is based on a $(p-1)$~-~fold degenerate vacuum structure, vacua can be associated with nodes of the ${\cal A}_{p-1}$ Dynkin diagram. The quasi particles in the scattering theory are kinks that interpolate neighboring vacua, they can be denoted by non-commutative symbols $K_{ab}(u)$ where $|a-b|=1$ with $a,b=1,\ldots,p-1$ and $u$ is related to the the kink rapidity $\theta$ by $u=-i\theta/p$, so that the physical strip is given by $0<{\rm Re}\;u<\pi/p$. In the rest of the paper, we will refer to $a,b$ as heights or spins. Formally, scattering between two kinks can be represented by the following equation (see Fig.(\ref{f1})) \begin{equation} K_{da}(u)K_{ab}(u^{'})=\sum_{c}S^{ab}_{dc}(u-u^{'}) K_{dc}(u^{'})K_{cb}(u) \end{equation} where the $S$-matrix is given by \begin{equation} S^{ab}_{dc}(u)={\cal U}(u) \left(\frac{[a][c]}{[d][b]}\right)^{-u/2\gamma}W^{ab}_{dc}(u) \label{eq:bulk} \end{equation} and \begin{equation} W^{ab}_{dc}(u)= \left(\sin u \delta_{bd}\left(\frac{[a][c]}{[d][b]}\right)^{1/2}+ \sin(\gamma-u)\delta_{ac}\right) \end{equation} satisfies the YBE in the RSOS representation. \begin{figure}[htbp] \centering \input{face1} \caption{\label{f1} The bulk RSOS scattering matrix $S^{ab}_{dc}(u-u^{'})$} \end{figure} Here $[a]$ denotes the usual $q$-number given by \[ [a]=\frac{\sin(a\gamma)}{\sin\gamma}\qquad \gamma=\frac{\pi}{p} \] and the overall factor ${\cal U}(u)$ is a product of Gamma functions satisfying the relations \begin{eqnarray*} {\cal U}(u){\cal U}(-u)\sin(\gamma-u) \sin(\gamma+u)&=&1\\ {\cal U}(\gamma-u)&=&{\cal U}(u)\;, \end{eqnarray*} and is given by \begin{eqnarray} {\cal U}(u)&=&\frac{1}{\pi}\Gamma\left(\frac{\gamma}{\pi}\right) \Gamma\left(1-{u\over{\pi}}\right) \Gamma\left(1-{\gamma\over{\pi}}+{u\over{\pi}}\right) \prod^{\infty}_{l=1}{F_l(u)F_l(\gamma-u)\over{F_l(0)F_l(\gamma)}},\nonumber\\ F_l(u)&=&{\Gamma\left({2l\gamma\over{\pi}}-{u\over{\pi}}\right) \Gamma\left(1+{2l\gamma\over{\pi}}-{u\over{\pi}}\right)\over{ \Gamma\left({(2l+1)\gamma\over{\pi}}-{u\over{\pi}}\right) \Gamma\left(1+{(2l-1)\gamma\over{\pi}}-{u\over{\pi}}\right)}}\;.\label{eq:fac} \end{eqnarray} This factor, together with the overall $q$-number factor, ensures that the $S$-matrix satisfies both crossing and unitarity constraints: \begin{eqnarray} S^{ab}_{dc}(u)&=&S^{bc}_{ad}(\gamma-u)\\ \sum_{c^{'}}S^{ab}_{dc^{'}}(u)S^{c^{'}b}_{dc}(-u)&=&\delta_{ac}\;. \end{eqnarray} Consider now the above scattering theory in the presence of a boundary denoted formally by ${\bf B}_{a}$, then the scattering between the kink and the boundary is described by the equation \begin{equation} K_{ab}(u){\bf B}_{a}=\sum_c R^b_{ac}(u) K_{bc}(-u){\bf B}_{c} \end{equation} which can be given a graphical representation shown in Fig.(\ref{f2}). Notice that in this representation, the boundary naturally carries an RSOS spin. \begin{figure}[htbp] \centering \input{face2} \caption{\label{f2} The boundary RSOS scattering matrix $R^b_{ac}(u)$} \end{figure} The function $R^{b}_{ac}$ is called the boundary scattering matrix and satisfies the BYBE, which in the RSOS representation takes the form \begin{eqnarray} \lefteqn{\sum_{a^{'},b^{'}}R^a_{bb^{'}}(u)S^{ac}_{b^{'}a^{'}}(u^{'}+u) R^{a'}_{b^{'}b^{''}}(u^{'})S^{a^{'}c}_{b^{''}a^{''}}(u^{'}-u)=} \nonumber\\ && \sum_{a^{'},b^{'}}S^{ac}_{ba^{'}}(u^{'}-u) R^{a^{'}}_{bb^{'}}(u^{'})S^{a^{'}c}_{b^{'}a^{''}}(u^{'}+u) R^{a''}_{b^{'}b^{''}}(u)\;. \label{eq:bybe} \end{eqnarray} This equation is illustrated graphically in Fig.(\ref{f}). \begin{figure}[htbp] \centering \input{face3} \caption{\label{f} The boundary Yang-Baxter equation} \end{figure} In general, the function $R^{a}_{bc}(u)$ can be written as \begin{equation} R^{a}_{bc}(u)={\cal R}(u)\left(\frac{[b][c]}{[a][a]}\right)^{-u/2\gamma} \left[\delta_{b\neq c}X^{a}_{bc}(u)+\delta_{bc} \left\{\delta_{b,a+1}U_a(u)+\delta_{b,a-1}D_a(u)\right\}\right]\label {eq:boundary} \end{equation} where ${\cal R}(u)$ has to be determined from the boundary crossing and unitarity constraints, while $X^{a}_{bc}$ and $U_a,D_a$ have to be determined from the BYBE. An overall $q$-number factor has also been multiplied to the above to cancel that from the bulk $S$-matrix in order to simplify the BYBE. If $X^{a}_{bc}$ does not vanish, the boundary $R$-matrix describes non-diagonal scattering process, otherwise the scattering is called diagonal. Note that due to the restriction that vacuum assumes value $1,\ldots,p-1$, $X^{1}_{bc},X^{p-1}_{bc},D_1,U_{p-1}$ are not defined. The case $p=3$ has only diagonal scattering, so $X^a_{bc}$ does not exist. \subsection{Non-diagonal scattering} We consider the scattering where the off-diagonal component $X^a_{bc}$ is non-vanishing. To start, consider the case $b\neq c\neq b^{''}$ in eqn.(\ref{eq:bybe}) where the BYBE gives \begin{equation} X^a_{a-1,a+1}(u^{'})X^{a+2}_{a+1,a+3}(u)=X^a_{a-1,a+1}(u) X^{a+2}_{a+1,a+3}(u^{'})\;;2\leq a\leq p-4 \end{equation} which implies that $X^{a}_{a\pm 1,a\mp 1}$ can be written as \begin{equation} X^{a}_{a\pm 1,a\mp 1}(u)=h_{\pm}(u)X_{\pm}^{a} \end{equation} where $h_{\pm}(u)$ depends only on $u$ and $X^{a}_{\pm}$ only on $a$. On the other hand, the case $c=b=b^{''}, a=a^{''}$ gives \begin{equation} X^a_{a-1,a+1}(u^{'})X^{a}_{a+1,a-1}(u)=X^a_{a-1,a+1}(u) X^{a}_{a+1,a-1}(u^{'})\;;2\leq a\leq p-2 \end{equation} which implies that \begin{equation} h_{+}(u^{'})h_{-}(u)=h_{+}(u)h_{-}(u^{'})\;, \end{equation} from which we conclude that \begin{equation} h_{+}(u)={\rm (const.)}h_{-}(u)\;. \end{equation} Absorbing the constant in the above equation into $X^{a}_{-}$ or $X^a_{+}$, we can make $h_{+}$ equal to $h_{-}$ so that we can absorb the $h_{\pm}(u)$ into the overall ${\cal R}(u)$ factor and treat $X^{a}_{bc}$ as $u$ independent from now on. With this simplification, eqn.(\ref{eq:bybe}) can be broken down into the following independent equations in addition to the above two equations: \begin{eqnarray} \lefteqn{U^{'}_{a}D_{a+2}f_{+}\left(1+f_{-}{[a]\over{[a+1]}}\right)+ D^{'}_{a+2}D_{a+2}f_{-}\left(1+f_{+}{[a+2]\over{[a+1]}}\right)} \nonumber \\ &&\mbox{}+X^{a+2}_{a+1,a+3}X^{a+2}_{a+3,a+1}f_{-}=U_{a}D^{'}_{a+2}f_{+} \left(1+f_{-}{[a+2]\over{[a+1]}}\right) \nonumber \\ &&\mbox{}+U^{'}_{a}U_{a}f_{-}\left(1+f_{+}{[a]\over{[a+1]}}\right) +X^{a}_{a-1,a+1}X^{a}_{a+1,a-1}f_{-}\label{eq:s0} \end{eqnarray} for $1\leq a\leq p-3$, \begin{eqnarray} &&D^{'}_{a+1}f_{-}\left(1+f_{+}{[a+1]\over{[a]}}\right)+ U^{'}_{a-1}f_{+}\left(1+f_{-}{[a-1]\over{[a]}}\right)\nonumber\\ &&\mbox{\hspace{1cm}}=U_{a-1}f_{+}-U_{a+1}f_{-}\label{eq:s1a}\\ &&\nonumber\\ &&U^{'}_{a}f_{-}\left(1+f_{+}{[a]\over{[a+1]}}\right)+ D^{'}_{a+2}f_{+}\left(1+f_{-}{[a+2]\over{[a+1]}}\right)\nonumber\\ &&\mbox{\hspace{1cm}}=D_{a+2}f_{+}-D_{a}f_{-}\label{eq:s1b} \end{eqnarray} for $2\leq a\leq p-3$, and \begin{eqnarray} &&U^{'}_{a-2}f_{+}f_{-}{[a][a-2]\over{[a-1]^2}}-U^{'}_{a} +D^{'}_{a}\left(1+f_{-}{[a]\over{[a-1]}}\right) \left(1+f_{+}{[a]\over{[a-1]}}\right)\nonumber\\ &&\mbox{\hspace{1cm}}=D_{a}\left(1+f_{+}{[a]\over{[a-1]}}\right) -U_{a}\left(1+f_{-}{[a]\over{[a-1]}}\right) \label{eq:s2a}\\&&\nonumber \\ &&D^{'}_{a+2}f_{+}f_{-}{[a][a+2]\over{[a+1]^2}}-D^{'}_{a}+ U^{'}_{a}\left(1+f_{-}{[a]\over{[a+1]}}\right) \left(1+f_{+}{[a]\over{[a+1]}}\right)\nonumber\\ &&\mbox{\hspace{1cm}}=U_{a}\left(1+f_{+}{[a]\over{[a+1]}}\right) -D_{a}\left(1+f_{-}{[a]\over{[a+1]}}\right)\label{eq:s2b} \end{eqnarray} for $2\leq a\leq p-2$. The last four equations are derived based on the assumption that the off-diagonal weight $X^a_{bc}$ is {\bf nonvanishing}. In the above equations, we used a compact notation where $U_a=U_a(u),\ U^{'}_a=U_a(u^{'})$ (similarly for $D_a$) and \[f_{\pm}=\sin(u^{'}\pm u)/\sin(\gamma-u^{'}\mp u)\;.\] In addition, it should also be mentioned that the last term in the rhs (lhs) of eqn.(\ref{eq:s0}) is present only when $a\neq 1 (p-3)$ and the first terms of eqns.(\ref{eq:s2a}) and (\ref{eq:s2b}) are allowed only for $a\neq 2$ and $a\neq p-2$, respectively. Let us call these terms that are not supposed to be there the ``unwanted'' terms. The way to solve these equations is to construct some recursion relations for the functions $U_a,D_a,X^a_{a\pm 1,a\mp 1}$ and solve the recursion relations subject to the conditions that the above mentioned ``unwanted'' terms are zero. Explicitly, one has \begin{eqnarray} X^{1}_{0,2}X^{1}_{2,0}&=X^{p-1}_{p-2,p}X^{p-1}_{p,p-2}&=0\;,\label{eq:bc}\\ U_0&=D_{p}&=0\;.\label{eq:bc2} \end{eqnarray} A few comments are in order here. Notice that the coefficients of $U_0$ and $D_p$ in eqns.(\ref{eq:s2a}), (\ref{eq:s2b}) are $[0]$ and $[p]$ respectively, which vanish by construction. So in principle, one needs not impose the above condition, eqn.(\ref{eq:bc2}), for the recursion relations of $U_a,D_a$. However, we shall see later that a particular solution of $U_a,D_a$ is given by \begin{equation} U_a=-D_a={\rm (const.)}\frac{1}{[a]} \end{equation} which cancels the vanishing coefficients at $a=0, p$, and renders the unwanted terms nonvanishing. Therefore, we have to impose eqn.(\ref{eq:bc2}) on the above solution. Notice also that most of the above equations do not apply to the case $p=4$, so we shall deal with this case separately. {}From the above it is clear that eqns.(\ref{eq:s0})-(\ref{eq:bc}) can be divided into two parts; eqn.(\ref{eq:s1b})-eqn.(\ref{eq:s2b}) determine $U_a$ and $D_a$ while eqns.(\ref{eq:s0}),(\ref{eq:bc}) determine $X^{a}_{bc}$. Indeed, comparing eqn.(\ref{eq:s1a}) with eqn.(\ref{eq:s2a}) and similarly eqn.(\ref{eq:s1b}) with eqn.(\ref{eq:s2b}), we deduce that \begin{eqnarray} (D_a^{'}-D_a)\left(1+f_{+}\frac{[a]}{[a-1]}\right)&=&(U^{'}_{a-2}-U_{a-2}) f_{+}\frac{[a]}{[a-1]}\nonumber\\ &&\mbox{}+(U_a^{'}-U_a)\\ (U_a^{'}-U_a)\left(1+f_{+}\frac{[a]}{[a+1]}\right)&=&(D^{'}_{a+2}-D_{a+2}) f_{+}\frac{[a]}{[a+1]}\nonumber\\ &&\mbox{}+(D_a^{'}-D_a)\;. \end{eqnarray} Substituting one into another, we get \begin{eqnarray} \frac{(U^{'}_{a+2}-U_{a+2})-(U^{'}_{a}-U_{a})}{(U^{'}_{a}-U_{a}) -(U^{'}_{a-2}-U_{a-2})}&=&\frac{\sin\left((a+1)\gamma+u^{'}+u\right)} {\sin\left((a-1)\gamma+u^{'}+u\right)}\label{eq:rec1}\\ \frac{(D^{'}_{a+2}-D_{a+2})-(D^{'}_{a}-D_{a})}{(D^{'}_{a}-D_{a}) -(D^{'}_{a-2}-D_{a-2})}&=&\frac{\sin\left((a+1)\gamma-u^{'}-u\right)} {\sin\left((a-1)\gamma-u^{'}-u\right)}\label{eq:rec2} \end{eqnarray} and writing the rhs respectively as \begin{eqnarray*} &\frac{\textstyle\cos\left(2u^{'}+(a+1)\gamma\right) -\cos\left(2u+(a+1)\gamma\right)} {\textstyle\cos\left(2u^{'}+(a-1)\gamma\right) -\cos\left(2u+(a-1)\gamma\right)}&\\ &\frac{\textstyle\cos\left(2u^{'}-(a+1)\gamma\right) -\cos\left(2u-(a+1)\gamma\right)} {\textstyle\cos\left(2u^{'}-(a-1)\gamma\right) -\cos\left(2u-(a-1)\gamma\right)}&\;, \end{eqnarray*} it is clear that \begin{eqnarray*} U_{a+2}(u)-U_{a}(u)&=&\cos\left(2u+(a+1)\gamma\right)+\beta^{'}_a\\ D_{a+2}(u)-D_{a}(u)&=&-\cos\left(2u-(a+1)\gamma\right)+\phi^{'}_a \end{eqnarray*} where $\beta^{'}_a,\phi^{'}_a$ are unknown functions of $a$ only. Iterating the above, one finds \begin{eqnarray} U_a(u)&\propto&\sin(2u+a\gamma)+\alpha(u)+\beta_a\\ D_a(u)&\propto&\sin(2u-a\gamma)+\gamma(u)+\phi_a \end{eqnarray} where $\alpha(u),\gamma(u)$ are unknown functions of $u$, and $\beta_a,\phi_a$ of $a$. Furthermore, from eqns.(\ref{eq:s0})-(\ref{eq:s2b}), one can establish the following symmetry \begin{equation} U_a(u)=-D_a(-u)\;,\label{eq:sym} \end{equation} which reduces the number of unknown functions to two, namely, $\alpha(u)$ and $\beta_a$. To determine them, we have to substitute the above expressions for $U_a,D_a$ back into eqns.(\ref{eq:s1a})-(\ref{eq:s2b}). Notice, however, that these equations are linear in $U_a,D_a$, so it suffices to consider $\alpha(u)$ and $\beta_a$ separately. Doing this amounts to finding special solutions to eqns.(\ref{eq:s1a})-(\ref{eq:s2b}) where $U_a,D_a$ have only $u$ or $a$ dependence. The solutions are given by \begin{equation} \alpha(u)={\rm(const.)}\frac{1}{\sin(2u)}\;,\hspace{2cm}\beta_a={\rm(const.)}\frac{1}{\sin(a\gamma)} \end{equation} respectively. Imposing eqn.(\ref{eq:bc2}) on $\alpha(u),\beta_a$, one finds \begin{equation} \beta_a=0\;. \end{equation} So far the above $U_a,D_a$ are obtained based on the assumption that the numerators and denominators of the lhs of eqns.(\ref{eq:rec1}),(\ref{eq:rec2}) are nonvanishing. In fact, there are additional solution to these recursion relations where the above mentioned numerators and denominators vanish; \begin{eqnarray*} U_{a+2}(u^{'})-U_{a}(u^{'})-U_{a+2}(u)+U_{a}(u)&=&0\\ D_{a+2}(u^{'})-D_{a}(u^{'})-D_{a+2}(u)+D_{a}(u)&=&0\;. \end{eqnarray*} The additional solution is given by \begin{equation} U_a(u)={\rm(const.)}\frac{\sin(a\gamma+2u)}{\sin(2u)\sin(a\gamma)}\;, \hspace{1cm} D_a(u)={\rm(const.)}\frac{\sin(a\gamma-2u)}{\sin(2u)\sin(a\gamma)}\;. \end{equation} However, on imposing the condition eqn.(\ref{eq:bc2}), one finds that they have to be zero. In summary, the general non-diagonal solution to the four linear equations is given by \begin{equation} \begin{array}{lcl} U_a(u)&=&\sin(2u+a\gamma)+{\rm (const.)}\frac{\textstyle 1} {\textstyle \sin(2u)} \\ D_{a+1}(u)&=&\sin\left(2u-(a+1)\gamma\right)+{\rm (const.)}\frac{\textstyle 1} {\textstyle \sin(2u)} \end{array}\label{eq:udn} \end{equation} where $1\leq a\leq p-2$. Having found $U_a,D_a$, the function $X^a_{bc}$ can be easily obtained from eqn.(\ref{eq:s0}), which can be further simplified with the symmetry properties given in eqn.(\ref{eq:sym}) and taking $u^{'}$ to be $-u$ since $X^a_{bc}$ does not depend on the rapidity. This gives \[X^{a+2}_{a+1,a+3}X^{a+2}_{a+3,a+1}-X^{a}_{a-1,a+1}X^{a}_{a+1,a-1} =D_{a+2}(u)U_{a+2}(u)-D_{a}(u)U_a(u)\;. \] Substituting $U_a,D_a$ into the rhs and iterating the equations, we get \begin{equation} X^{a}_{a-1,a+1}X^{a}_{a+1,a-1}=\sin^2\gamma-\sin^2(a\gamma)\label{eq:xn} \end{equation} where the use of eqn.(\ref{eq:bc}) forces the first term to be $\sin^2\gamma$ and the coefficient of $\frac{1}{\sin(2u)}$ to vanish. Since this equation determines only the product, $X^{a}_{a-1,a+1}$ and $X^{a}_{a+1,a-1}$ are determined up to a gauge factor. These solutions have the property that \begin{equation} U_{p-a}(u)=-D_a(u)\;,\hspace{1cm}X^{p-a}_{p-a-1,p-a+1} X^{p-a}_{p-a+1,p-a-1}=X^{a}_{a-1,a+1}X^{a}_{a+1,a-1}\;. \end{equation} Next, we consider the BYBE for $p=4$. The functions $U_2,D_2$ and $X^2_{13},X^2_{31}$ satisfy the following equations \begin{eqnarray} X^{2}_{13}(u^{'})X^2_{31}(u)&=&X^{2}_{13}(u)X^{2}_{31}(u^{'})\label{eq:ss0}\\ \lefteqn{U_{2}\left(1+\sqrt{2}f_{-}\right)+D_{2}^{'}\left(1+\sqrt{2} f_{+}\right)\left(1+\sqrt{2}f_{-}\right)}\hspace{3cm}\nonumber\\ &=&U_2^{'}+D_{2}\left(1+\sqrt{2}f_{+}\right)\\ \lefteqn{D_{2}\left(1+\sqrt{2}f_{-}\right)+U_{2}^{'}\left(1+\sqrt{2} f_{+}\right)\left(1+\sqrt{2}f_{-}\right)}\hspace{3cm}\nonumber\\ &=&D_2^{'}+U_{2}\left(1+\sqrt{2}f_{+}\right)\;. \end{eqnarray} The functions $U_1, D_3$ are diagonal scattering components and do not couple to the above equations and we shall defer to next section for their computation. Here we have used the compact notation introduced earlier for $U_a, D_a$, and written out explicitly the rapidity dependence of $X^a_{bc}$. As before, the last two equations have been derived based on the assumption that $X^2_{13},X^2_{31}$ are nonvanishing. From eqn.(\ref{eq:ss0}), we deduce that \begin{equation} X^2_{13}(u)={\rm (const.)}X^2_{31}(u)\;. \end{equation} We are free to take $X^2_{31}$ as unity and the above implies that $X^2_{13}$ is just a gauge factor, which we call $g$. The rest of the equations can be turned into ordinary differential equations in the limit $u^{'}\rightarrow u$, giving \begin{eqnarray} \left(\dot{U}_2(u)+\dot{D}_2(u)\right)\tan(2u)+2\left(U_2(u) +D_2(u)\right)&=&0\\ \left(\dot{U}_2(u)-\dot{D}_2(u)\right)\cot(2u)-2\left(U_2(u) -D_2(u)\right)&=&0\;, \end{eqnarray} which can be integrated to give \begin{eqnarray} U_2(u)&=&B/\sin(2u)+C\cos(2u)\\ D_2(u)&=&B/\sin(2u)-C\cos(2u)\;, \end{eqnarray} with $B,C$ as free parameters. This completes the determination of the non-diagonal solutions of the BYBE. \subsection{Diagonal scattering} For the diagonal scattering, we take \begin{equation} R^{a}_{bc}(u)=([b]/[a])^{-u/\gamma}{\cal R}(u) \delta_{bc}\left[\delta_{b,a+1}U_a(u)+\delta_{b,a-1}D_a(u)\right]\;, \end{equation} and the BYBE is equivalent to a single equation \begin{eqnarray} \lefteqn{U_{a-1}(u^{'})D_{a+1}(u)\sin(u^{'}+u) \sin(a\gamma-u^{'}+u)}\nonumber\\ &&\mbox{}+D_{a+1}(u^{'})D_{a+1}(u)\sin(u^{'}-u) \sin(a\gamma+u^{'}+u)=\nonumber\\ \lefteqn{U_{a-1}(u)D_{a+1}(u^{'})\sin(u^{'}+u) \sin(a\gamma+u^{'}-u)}\nonumber\\ &&\mbox{}+U_{a-1}(u^{'})U_{a-1}(u)\sin(u^{'}-u) \sin(a\gamma-u^{'}-u)\;,\label{eq:dia} \end{eqnarray} which holds only for $2\leq a\leq p-2$. So the functions $U_{p-2}$ and $D_2$ can not be determined from the BYBE. The above can be recast into \begin{eqnarray} \lefteqn{[\cos(a\gamma+2u)Z(u)-\cos(a\gamma-2u)][Z(u^{'})-1]=}\nonumber\\ &&\mbox{ }[\cos(a\gamma+2u^{'})Z(u^{'})-\cos(a\gamma-2u^{'})][Z(u)-1] \end{eqnarray} where \[Z_a(u)\equiv D_{a+1}(u)/U_{a-1}(u)\;.\] One can easily find that the general solution is \begin{equation} \frac{D_{a+1}(u)}{U_{a-1}(u)}=\frac{\sin(\xi_a+u)\sin(\xi_a+a\gamma-u)} {\sin(\xi_a-u)\sin(\xi_a+a\gamma+u)}\label{eq:ratio} \end{equation} where $\xi_a$ is a free parameter. Thus for the diagonal solution, there are $p-3$ parameters $\xi_a$. This solution gives $p-1$ distinct diagonal scattering theories of kinks, each with a specific boundary ${\bf B}_a$. There is one free parameter $\xi_a$ for each theory, except for the cases of $a=1,p-1$ where there is no free parameter. This solution includes a particular case of $p=4$ which has been omitted earlier. Further relations from boundary unitarity and crossing symmetry will be required to disentangle $U_{a-1}$ and $D_{a+1}$, and determine $U_2$ and $D_{p-2}$, see later. \subsection{Boundary unitarity and crossing symmetry} The boundary unitarity and crossing symmetry conditions of the scattering matrix $R^a_{bc}(u)$ determine to some extent the overall factor $R(u)$. These conditions can be written as \begin{eqnarray} \sum_{c}R^{a}_{bc}(u)R^{a}_{cd}(-u)&=&\delta_{bd}\\ \sum_{d}S^{ac}_{bd}(2u)R^{d}_{bc}(\gamma/2+u)&=&R^{a}_{bc}(\gamma/2-u)\;. \end{eqnarray} Consider the non-diagonal scattering ($p>4$) first. Substituting the expression for $R^a_{bc}$ into the unitarity condition, we get the following \[{\cal R}(u){\cal R}(-u)\left[X^a_{bd}X^a_{db}\delta_{b\neq d} -U_a(u)D_a(u)\right]=1\] where use has been made of the symmetry eqn.(\ref{eq:sym}). Applying the results eqns.(\ref{eq:udn}), (\ref{eq:xn}) to the above leads to \begin{equation} {\cal R}(u){\cal R}(-u)\left(\sin^2\gamma-4\sin^2u+4\sin^4u\right)=1\;. \label{eq:r1} \end{equation} While for crossing symmetry condition we get \begin{equation} {\cal U}(2u){\cal R}(\gamma/2+u)\sin(\gamma-2u)= {\cal R}(\gamma/2-u)\label{eq:r2} \end{equation} using the relations \begin{eqnarray} &&D_{a+2}(\gamma-u){[a+2]\over{[a+1]}}+ U_a(\gamma-u){[a]\over{[a+1]}}\nonumber\\ &&\hspace{2cm}=f(2u)\left(U_a(u)-U_a(\gamma-u)\right)\\ &&U_a(\gamma-u){[a]\over{[a+1]}}+D_{a+2}(\gamma-u) {[a+2]\over{[a+1]}}\nonumber\\ &&\hspace{2cm}=f(2u)\left(D_{a+2}(u)-D_{a+2}(\gamma-u)\right) \end{eqnarray} which are obtained from eqn.(\ref{eq:s1b}) in the limit $u^{'}\rightarrow \gamma-u$. Here $f(u)=\sin u /\sin(\gamma-u)$. The factor ${\cal R}(u)$ can be determined from eqns.(\ref{eq:r1}), (\ref{eq:r2}) up to the usual CDD ambiguity by separating ${\cal R}(u)={\cal R}_0(u){\cal R}_1(u)$ where ${\cal R}_0$ satisfies \begin{equation} \begin{array}{lll} {\cal R}_{0}(u){\cal R}_0(-u)&=&1\\ {\cal U}(2u){\cal R}_0(\gamma/2+u)\sin(\gamma-2u)&=&{\cal R}_0(\gamma/2-u)\;, \end{array} \end{equation} whose minimal solution reads \begin{equation} R_0(u)=\frac{F_0(u)}{F_0(-u)} \end{equation} where $F_l(u)$ has been given in eqn.(\ref{eq:fac}). While ${\cal R}_1$ satisfies \begin{equation} \begin{array}{ll} &{\cal R}_1(u){\cal R}_1(-u)\left(\sin^2\gamma-4\sin^2u+4\sin^4u\right)=1\\ &{\cal R}_1(u)={\cal R}_1(\gamma-u) \end{array} \end{equation} with minimum solution \begin{equation} {\cal R}_1(u)=\frac{1}{2}\sigma(\frac{\gamma}{2},u)\sigma(\frac{\pi-\gamma}{2},u) \end{equation} where \begin{equation} \begin{array}{lcl} \sigma(x,u)&=&\frac{\textstyle\prod(x,\frac{\gamma}{2}-u) \prod(-x,\frac{\gamma}{2}-u) \prod(x,-\frac{\gamma}{2}+u)\prod(-x,-\frac{\gamma}{2}+u)} {\textstyle\prod^2(x,\frac{\gamma}{2})\prod^2(-x,\frac{\gamma}{2})}\\&&\\ \prod(x,u)&=&\displaystyle\prod_{l=0}^{\infty} \frac{\textstyle\Gamma\left(\frac{1}{2}+(2l+\frac{1}{2})\frac{\gamma}{\pi} +\frac{x}{\pi}-\frac{u}{\pi}\right)} {\textstyle\Gamma\left(\frac{1}{2}+(2l+\frac{3}{2}) \frac{\gamma}{\pi}+\frac{x}{\pi}-\frac{u}{\pi}\right)}\;. \end{array} \end{equation} For $p=4$, the crossing symmetry condition is the same as eqn.(\ref{eq:r2}), but unitarity now requires that \begin{equation} {\cal R}(u){\cal R}(-u)\left(g+C^2\cos^2(2u)-B^2/\sin^2(2u)\right)=1 \end{equation} These equations can again be solved by separation (see \cite{meahn} for details). It should be remarked that there are two parameters in this case while there is only one in the higher $p$ cases. Next, we consider the diagonal case. For convenience, we set the undefined terms $U_{0},D_{p}$ to be zero. Unitarity and crossing symmetries relations give respectively \begin{equation} \begin{array}{lll} {\cal R}(u){\cal R}(-u)U_a(u) U_a(-u)&=&1\\ {\cal R}(u){\cal R}(-u)D_{a+1}(u) D_{a+1}(-u)&=&1 \end{array} \end{equation} and \begin{equation} \begin{array}{c} {\cal U}(\gamma-2u){\cal R}(\gamma-u) \left[U_a(\gamma-u)\sin\gamma\sin(a\gamma+2u)\right.\\ \mbox{ }\left.+D_{a+2}(\gamma-u)\sin(\gamma-2u) \sin\left((a+2)\gamma\right)\right] ={\cal R}(u)U_a(u)\sin((a+1)\gamma)\\ \\ {\cal U}(\gamma-2u){\cal R}(\gamma-u) \left[D_{a+1}(\gamma-u)\sin\gamma\sin\left((a+1)\gamma-2u\right)\right.\\ \mbox{ }\left.+U_{a-1}(\gamma-u)\sin(\gamma-2u) \sin\left((a-1)\gamma\right)\right] ={\cal R}(u)D_{a+1}(u)\sin(a\gamma) \end{array} \end{equation} for $ 1\leq a\leq p-2$. These equations can be solved separately; we set \begin{equation} \begin{array}{lll} {\cal R}(u){\cal R}(-u)&=&1\\ {\cal U}(2u){\cal R}(\gamma/2+u)\sin(\gamma-2u)&=&{\cal R}(\gamma/2-u)\;, \end{array} \end{equation} so that ${\cal R}(u)$ has exactly the same solution as that of ${\cal R}_0(u)$ considered earlier. While $U_a,D_a$ satisfy \begin{equation} \begin{array}{lll} U_a(u)U_a(-u)&=&1\\ D_{a+1}(u) D_{a+1}(-u)&=&1 \end{array}\;\label{eq:u1} \end{equation} and \begin{equation} \begin{array}{l} U_a(\gamma-u)\sin\gamma\sin(a\gamma+2u)+D_{a+2}(\gamma-u) \sin(\gamma-2u)\sin((a+2)\gamma)\\ \mbox{\hspace{2cm}}=U_a(u)\sin(2u)\sin((a+1)\gamma)\;,\\ D_{a+1}(\gamma-u)\sin\gamma\sin((a+1)\gamma-2u)+U_{a-1}(\gamma-u) \sin(\gamma-2u)\sin((a-1)\gamma)\\ \mbox{\hspace{2cm}}=D_{a+1}(u)\sin(2u)\sin(a\gamma)\;. \end{array} \end{equation} for $1\leq a\leq p-2$. Substituting eqn.(\ref{eq:ratio}) into the above we get a relation between $U_a(u)$ ($D_a(u)$) and $U_a(\gamma-u)$ ($D_a(\gamma-u)$); \begin{equation} \begin{array}{lll} \frac{\textstyle U_{a-1}(u)}{\textstyle U_{a-1}(\gamma-u)}&=& \frac{\textstyle \sin(2(\gamma-u))\sin(\xi_a-u)\sin(\xi_a+a\gamma+u) }{\textstyle \sin(2u)\sin(\xi_a-(\gamma-u))\sin(\xi_a+a\gamma+(\gamma-u))}\\ &&\\ \frac{\textstyle D_{a+1}(u)}{\textstyle D_{a+1}(\gamma-u)}&=& \frac{\textstyle \sin(2(\gamma-u))\sin(\xi_a+u) \sin(\xi_a+a\gamma-u)}{\textstyle \sin(2u) \sin(\xi_a+(\gamma-u))\sin(\xi_a+a\gamma-(\gamma-u))}\;. \end{array}\label{eq:diag2} \end{equation} These relations together with eqn.(\ref{eq:u1}) give the minimal solutions of $U_a(u)$ and $D_a(u)$ \begin{eqnarray} U_{a-1}(u)&=&\frac{1}{2}\sin 2(\gamma-u)\sin(\xi_a-u)\sin(\xi_a+a\gamma+u) \sigma(\gamma,u)\nonumber\\ &&\mbox{}\times\sigma(\frac{\pi}{2}-\gamma,u) \sigma(\frac{\pi}{2}-\xi_a,u)\sigma(\frac{\pi}{2}-\xi_a-a\gamma,u) \label{eq:udsol1}\;,\\ D_{a+1}(u)&=&\frac{1}{2}\sin 2(\gamma-u)\sin(\xi_a+u)\sin(\xi_a+a\gamma-u) \sigma(\gamma,u)\nonumber\\ &&\mbox{}\times\sigma(\frac{\pi}{2}-\gamma,u) \sigma(\frac{\pi}{2}-\xi_a,u)\sigma(\frac{\pi}{2}-\xi_a-a\gamma,u)\;. \label{eq:udsol2} \end{eqnarray} for $2\leq a\leq p-2$ and \begin{eqnarray} U_{p-2}(u)&=&\frac{1}{2}\sin 2(\gamma-u)\sigma(\gamma,u) \sigma(\frac{\pi}{2}-\gamma,u)\;,\\ D_{2}(u)&=&\frac{1}{2}\sin 2(\gamma-u)\sigma(\gamma,u) \sigma(\frac{\pi}{2}-\gamma,u)\;. \end{eqnarray} To summarize, there are two classes of solutions to the BYBE; diagonal and non-diagonal. Unlike the solution in vertex representation, the former is not a special limit of the latter. In fact, the diagonal solution carries $p-3$ free parameters while nondiagonal has non. \subsection{Comments on the SOS model} We have considered from the start heights take values from $1$ to $p-1$, which is necessary for the bulk scattering weights to be finite as the parameter $\pi/\gamma=p$ is a positive integer. When $\pi/\gamma$ is not a rational number, there is no bounds on the heights and the corresponding representation is known as solid-on-solid (SOS). The removal of the heights' restriction certainly affects the solution of the BYBE. For the diagonal solution, it is clear that essentially there is no difference between the SOS and RSOS solution, where the SOS solution is given by eqns.(\ref{eq:udsol1}),(\ref{eq:udsol2}) for any integer $a$ and the overall factor ${\cal R}(u)$ is the same as before. Hence there is a free parameter for each height. While for the non-diagonal solution, the conditions eqns.(\ref{eq:bc}),(\ref{eq:bc2}) of the recursion relations do not apply at all, thus $\beta_a\neq 0$ and the corresponding diagonal weights are given by \begin{equation} \begin{array}{lcl} U_a(u)&=&c_1\sin(2u+a\gamma)+c_2\frac{\textstyle 1}{\textstyle \sin(a\gamma)} +c_3\frac{\textstyle \sin(a\gamma+2u)}{\textstyle \sin(2u)\sin(a\gamma)} +c_4\frac{\textstyle 1}{\textstyle \sin(2u)}\\ \\ D_a(u)&=&c_1\sin(2u-a\gamma)-c_2\frac{\textstyle 1}{\textstyle \sin(a\gamma)} +c_3\frac{\textstyle \sin(a\gamma-2u)}{\textstyle \sin(2u)\sin(a\gamma)} +c_4\frac{\textstyle 1}{\textstyle \sin(2u)} \end{array} \end{equation} where $c_i\;;i=1,\ldots,4$ are free parameters. The off-diagonal weights satisfy \begin{eqnarray} X^{a}_{a-1,a+1}X^{a}_{a+1,a-1}&=&c^2_0+2c_1c_4\cos(a\gamma) -c_1^2\sin^2(a\gamma)-c_2^2\frac{1}{\sin^2(a\gamma)} \nonumber\\ &&\mbox{}-2c_2c_3\frac{\cos(a\gamma)}{\sin^2(a\gamma)} -c_3^2\frac{1}{\sin^2(a\gamma)} \label{eq:xn1}\; \end{eqnarray} and are determined up an additive constant $c_0$. Of these five free parameters one is an overall factor, so there are four free parameters in this case. The overall factor ${\cal R}_0$ is given as before, but ${\cal R}_1(u)$ now contains all the information of the boundary conditions and has to satisfy \begin{equation} \begin{array}{l} {\cal R}_1(u){\cal R}_1(-u)\left(c_0^2-c_1^2\sin^2(2u)+2c_1c_2\cos(2u)- \frac{\textstyle c_3^2+2c_3c_4\cos(2u)+c_4^2}{\textstyle\sin^2(2u)}\right)=1\\ {\cal R}_1(u)={\cal R}_1(\gamma-u) \end{array} \end{equation} whose minimum solution is \begin{equation} {\cal R}_1(u)=\frac{2\sigma(\vartheta_1,u)\sigma(\vartheta_2,u) \sigma(\vartheta_3,u)\sigma(\vartheta_4,u) }{(c_3+c_4)\sigma(0,u)}{\cal S}(u) \end{equation} where $\vartheta_i$ are related to $c_i$ via \begin{eqnarray*} \prod_{i=1}^4\cos^2\vartheta_1&=&\left(\frac{c_3+c_4}{4c_1}\right)^2\;,\\ \sum_{i<j<k=1}^4\cos^2\vartheta_i\cos^2\vartheta_j\cos^2\vartheta_k &=&\frac{c_0^2+c_3c_4+2c_1c_2}{4c_1^2}\;,\\ \sum_{i<j=1}^4\cos^2\vartheta_i\cos^2\vartheta_j &=&\frac{c_0^2+5c_1c_2+4c_1^2}{4c_1^2}\;,\\ \sum_{i=1}^4\cos^2\vartheta_i&=&\frac{c_1c_2+2c_1^2}{c_1^2} \end{eqnarray*} and $\sigma(x,u)$ has been given before, while \begin{equation} {\cal S}(u)=\prod_{l=0}^{\infty}\frac{\Gamma\left((2l+1)\frac{\gamma}{\pi} +\frac{u}{\pi}\right)\Gamma\left((2l+2)\frac{\gamma}{\pi}+\frac{u}{\pi}\right) \Gamma\left(1-(2l+1)\frac{\gamma}{\pi}-\frac{u}{\pi}\right) \Gamma\left(1-(2l+2)\frac{\gamma}{\pi}-\frac{u}{\pi}\right)} {\Gamma\left((2l+1)\frac{\gamma}{\pi}-\frac{u}{\pi}\right) \Gamma\left((2l+2)\frac{\gamma}{\pi}-\frac{u}{\pi}\right) \Gamma\left(1-(2l+1)\frac{\gamma}{\pi}+\frac{u}{\pi}\right) \Gamma\left(1-(2l+2)\frac{\gamma}{\pi}+\frac{u}{\pi}\right)}\;. \end{equation} It is interesting to point out that the $c_2,c_3,c_4$ terms in $U_a,D_a$ are related to the boundary $K$-matrix of the 6-vertex model\cite{gon} via the vertex-SOS intertwiner\cite{bax,zou,men}. In this light, one can regard the term $c_1$ to be the solution special to the SOS/RSOS representation of the BYBE. In fact, it is not clear whether there exist an intertwiner that relate it to the $K$-matrix of the 6-vertex model. It is then natural to wonder if similar special solution occur in the elliptic case\cite{hou}. Also intriguing is that restricting SOS to RSOS corresponds to \begin{equation} c_2=c_3=c_4=0,\qquad c_0=c_1\sin(a\gamma)\;. \end{equation} \section{Commuting transfer matrix} Following the technique proposed in \cite{skly} for the vertex model, one can similarly construct a family of commuting transfer matrix for the RSOS/SOS model with boundary. To start, it can be shown that if $R^{a}_{bc}$ is a solution to the BYBE in the RSOS/SOS form then the following combination \begin{equation} \sum_a S^{fe}_{ba}(u-u_1)S^{ae}_{cd}(u+u_1)R^a_{bc}(u) \end{equation} also satisfies the BYBE, where $S^{fe}_{ba}(u)$ is the solution of the bulk YBE given in eqn.(\ref{eq:bulk}) and $u_1$ is an arbitrary parameter. The proof is essentially the same as the vertex case given in \cite{skly} and we shall not repeat it here. It is convenient to think of the BYBE as the defining relation of some associative algebra generated by the symbol $R^a_{bc}$. So the solutions given in the previous section correspond to particular representations of this algebra where the ``quantum space'' is trivial and the auxiliary space is the space of a one-step path ${\cal P}_1$ on a truncated Bratteli diagram with $ab$ and $ac$ being in- and out-state, respectively. In this context, the above ``decorated'' solution then corresponds to a representation whose quantum space is isomorphic to ${\cal P}_1$ that is formed by the nodes $f,b$ ($d,c$). Clearly, the above construction can be repeated for an arbitrary number of times (say $N+1$) giving a boundary $R$-matrix that acts on ${\cal P}_{N}$, the collection of $N$-step paths on a truncated Bratteli diagram. We shall denote such a solution as ${\bf R}^{a}_{bc}$ which should be regarded as an operator acting on ${\cal P}_{N}$. Its matrix element is explicitly given by \begin{eqnarray} {\bf R}^a_{bc}(u)_{a_0,\cdots,a_{N};a^{''}_0,\cdots, a^{''}_{N}}&=&\delta_{aa^{'}_N}\delta_{b a_N} \delta_{ca^{''}_N}\prod_{i=1}^N\sum_{a^{'}_{i-1}} \left( S^{a_ia^{'}_i} _{a_{i-1}a^{'}_{i-1}}(u-u_i)\right.\nonumber\\ &&{}\times\left.S^{a^{'}_{i-1}a^{'}_i}_{a^{''}_{i-1} a^{''}_{i}}(u+u_i)\right)R^{a^{'}_{0}}_{a_{0} a^{''}_{0}}(u)\label{eq:matrix} \end{eqnarray} which has the graphical representation given in Fig.(\ref{f4}). \begin{figure}[htbp] \centering \input{face4} \caption{\label{f4} ``Decorated" boundary scattering matrix} \end{figure} It carries $N$ parameters $u_i$ from the bulk $S$-matrices and additional ones from $R^a_{bc}$. To form a commuting transfer matrix out of ${\bf R}^a_{bc}$, like in the vertex case, one has to combine it with another BYBE solution (denoted here as $\tilde{R}$) as follows \begin{equation} {\bf T}(u)_{a_0,\cdots,a_{N};a^{''}_0,\cdots, a^{''}_N}\equiv \sum_{a,b,c}\tilde{R}^{a}_{cb}(u) {\bf R}^a_{b,c}(u)_{a_0,\cdots, a_N;a^{''}_0,\cdots,a^{''}_N}\;. \label{eq:transfer} \end{equation} Hence the transfer matrix ${\bf T}(u)$ is again an operator acting on ${\cal P}_{N}$. To show that \[[{\bf T}(u),{\bf T}(u^{'})]=0\;,\] one inserts four bulk $S$-matrices using the unitarity condition \[\sum_{\alpha}S^{ac}_{b\alpha}(u^{'}+u) S^{\alpha c}_{bd}(-u^{'}-u)\propto \delta_{ac}\;,\] and the crossing-unitarity condition \[\sum_{\alpha}S^{b\alpha}_{cd}(u^{'}+u) S^{d\alpha}_{ab}(\gamma-u+\gamma-u^{'})\propto \delta_{ac}\;,\] into ${\bf T}(u){\bf T}(u^{'})$. Then, one uses the BYBE to permute the ${\bf R}$'s, and the $\tilde{R}$'s. Because of the argument $\gamma-u+\gamma-u^{'}$ that appears in the crossing-unitarity condition, one can take \begin{equation} \tilde{R}^{a}_{bc}(u)\equiv R^a_{bc}(\gamma-u)\label{eq:rbm}\;. \end{equation} \section{Diagonalization of the Transfer matrix} So far, we managed to obtain solutions to the BYBE and construct the corresponding commuting transfer matrix. It would be necessary to diagonalize the transfer matrix in order to study the statistical models given by these solutions. For applications to field theory, diagonalization of the transfer matrix is also needed in order to write down the thermodynamic Bethe ansatz equation. For this purpose, a systematic approach generalizing the algebraic Bethe ansatz for the case of the periodic boundary condition has been devised in \cite{skly}. However, the method relies upon the conservation of the $S^z$ in the vertex language and is thus applicable only to diagonal boundary scattering theories. Therefore, we shall consider only the diagonal scattering solution and adapt the algebraic Bethe ansatz method devised in \cite{skly}, along the line of \cite{devg}, to the SOS model with boundary. The algebraic Bethe ansatz relies upon the existence of some pseudo-vacuum, which in the vertex model, is a state with either all spins equal to $1/2$ or $-1/2$. In the SOS model, such a state corresponds to a path which takes the form of a $45^{\rm o}$-oriented straight line in the Bratteli diagram. It is obvious that some heights on such a path, for lattice size $N$ large enough ($>p$), have to exceed the bounds $1$ or $p-1$. Therefore, this pseudo-vacuum does not belong to the {\bf truncated} Bratteli diagram and the algebraic Bethe ansatz that we are going to use is applicable to the SOS model only. For consistency, we have to assume that $\gamma/\pi$ is an irrational number. To diagonalize the transfer matrix given in eqn.(\ref{eq:transfer}), we have to first write down the algebraic relations satisfied by ${\bf R}^a_{bc}$. As before, we express the operator ${\bf R}^a_{bc}$ as follows \begin{eqnarray} {\bf R}^{a}_{bc}(u)&=&{\cal R}_N(u)\left(\frac{[b][c]}{[a][a]}\right)^{-u/2\gamma} \prod_{j=2}^{N}\left(\frac{[a_j^{''}]}{[a_j]}\right)^{(u_j-u_{j-1})/2\gamma} \left(\delta_{b\neq c}{\bf X}^{a}_{bc}(u)\right. \nonumber\\ &&\mbox{\hspace{2cm}}\left.+\delta_{bc} (\delta_{b,a+1}{\bf U}_a(u)+\delta_{b,a-1}{\bf D}_a(u))\right) \label{eq:operator} \end{eqnarray} where \[{\cal R}_N(u)={\cal R}(u)\prod_{i=1}^{N}{\cal U}(u-u_i){\cal U}(u+u_i)\;,\] which, along with $q$-number factors, ensures the boundary crossing and unitarity symmetry of ${\bf R}^a_{bc}$. Here ${\bf X}^a_{bc},{\bf U}_a,{\bf D}_a$ are now non-commutative operators that satisfy the algebraic relations encoded by the BYBE. (see appendix) Because $\tilde{R}^a_{bc}$ is diagonal, the commuting transfer matrix can be written as \begin{eqnarray} {\bf T}(u)&=&{\cal R}(\gamma-u){\cal R}_N(u)\prod_{j=1}^{N-1} \left(\frac{[a_j^{''}]}{[a_j]}\right)^{(u_j-u_{j+1})/2\gamma} \left(\frac{[b-1]}{[b]}U_{b-1}(\gamma-u){\bf U}_{b-1}(u)\right.\nonumber\\ &&\mbox{\hspace{2cm}}\left.+ \frac{[b+1]}{[b]}D_{b+1}(\gamma-u){\bf D}_{b+1}(u)\right)\;.\label{eq:comtran} \end{eqnarray} It is important to bear in mind that the RSOS heights $a_{0},a_{0}^{''}$ on the ``bottom'' boundary (see Fig.(\ref{f4})) are set to be the same since we consider only diagonal scattering. Let us denote them as $a$, while the heights on the other boundary are taken to be $b$ as evident from the above equation. Hence the statistical model associated with the transfer matrix is the one defined on a square lattice whose boundary heights are shown in Fig.(\ref{f5}), where heights denoted by open-circles are summed over. \begin{figure}[htbp] \centering \input{face5} \caption{\label{f5} Lattice generated by the transfer matrix } \end{figure} Consider the column of heights in Fig.(\ref{f4}) where $a_j=a+j$ and denote this state as $\omega_a^{a+N}$. By construction, $\omega_a^{a+N}$ is an eigenstate of ${\bf U}_{a+N-1}(u)$ and ${\bf D}_{a+N+1}(u)$ since the top and bottom heights of ${\bf U}_{a+N-1}(u)\omega_a^{a+N}$ and ${\bf D}_{a+N+1}(u)\omega_a^{a+N}$ are $a+N$ and $a$ respectively. Furthermore, it is annihilated by ${\bf X}^{a+N+1}_{a+N+2,a+N}(u)$ due to the constraint that neighboring heights differ by $\pm 1$. Explicitly, \begin{eqnarray} {\bf U}_{a+N-1}(u)\omega_a^{a+N}&=&{\cal U}_{a}^{N}(u)\omega_a^{a+N}\nonumber\\ {\bf D}_{a+N+1}(u)\omega_a^{a+N}&=&{\cal D}_{a}^{N}(u)\omega_a^{a+N}\nonumber\\ {\bf X}^{a+N+1}_{a+N+2,a+N}(u)\omega_a^{a+N}&=&0 \end{eqnarray} where the eigenvalues are given by \begin{eqnarray} {\cal U}_a^N(u)&=&\sum_{j=1}^{N}\left(\prod_{i=j+1}^{N}{\cal A}_{a+i}(u_i) \right){\cal B}_{a+j}(u_j)\left(\prod_{k=1}^{j-1}{\cal C}_{a+k}(u_k)\right) D_{a+1}(u)\nonumber\\ &&\mbox{}+\prod_{j=1}^{N}{\cal A}_{a+j}(u_j)U_{a-1}(u)\nonumber\\ {\cal D}_a^N(u)&=&\prod_{j=1}^{N}{\cal C}_{a+j}(u_i)D_{a+1}(u) \end{eqnarray} with \[\begin{array}{lll} {\cal A}_{a+i}(u_i)&\equiv&\frac{\textstyle \sin((a+i)\gamma) \sin((a+i-2)\gamma)\sin(u-u_i)\sin(u+u_i)} {\textstyle \sin^2((a+i-1)\gamma)}\\ {\cal B}_{a+i}(u_i)&\equiv&\frac{\textstyle\sin^2\gamma \sin((a+i-1)\gamma+u-u_i) \sin((a+i-1)\gamma+u+u_i)}{\textstyle\sin^2((a+i-1)\gamma)}\\ {\cal C}_{a+i}(u_i)&\equiv&\sin(\gamma-u+u_i)\sin(\gamma-u-u_i) \end{array}\] and $U_{a-1}(u),D_{a+1}(u)$ are given by eqns.(\ref{eq:udsol1}),(\ref{eq:udsol2}). The state ${\bf X}^{a+N-1}_{a+N-2,a+N}(\lambda)\omega_a^{a+N}$ is non-zero, which corresponds to a column of spins where the ``bottom'' and ``top'' spins have heights $a$ and $a+N-2$ respectively. So to obtain a state whose top spin has height given by $b$, one has to act on $\omega_a^{a+N}$ successively with $M\equiv (N+a-b)/2$ operators ${\bf X}^{a+N-i}_{a+N-i-1,a+N-i+1}(\lambda_i)\;\;,i=1,\cdots,M$. We shall denote such a state as \begin{equation} \psi_a^b(\vec{\lambda})\equiv {\bf X}^{b+1}_{b,b+2}(\lambda_1) \cdots{\bf X}^{a+N-1}_{a+N-2,a+N}(\lambda_M)\omega_a^{a+N}\label{eq:state} \end{equation} which is the Bethe ansatz state, and the set of parameters $\vec{\lambda}\equiv (\lambda_1,\cdots,\lambda_{M})$ have to satisfy some consistency condition necessary for $\psi_a^b(\vec{\lambda})$ to be an eigenstate of the transfer matrix. Notice that $M$ is always an integer fixed by the heights $a,b$ and $N$. The algebraic relations among ${\bf X}^a_{a\pm 1,a\mp 1}, {\bf U}_a,{\bf D}_a$ necessary for our purpose are (see Appendix) \begin{eqnarray} {\bf U}_{a-1}(u^{'}){\bf X}^{a+1}_{a,a+2}(u)&=&g_{1a}(u^{'},u) {\bf X}^{a+1}_{a,a+2}(u){\bf U}_{a+1}(u^{'}) +g_{2a}(u^{'},u){\bf X}^{a+1}_{a,a+2}(u^{'}){\bf U}_{a+1}(u)\nonumber\\ &+&g_{3a}(u^{'},u){\bf X}^{a+1}_{a,a+2}(u^{'}){\bf D}_{a+3}(u) +g_{4a}(u^{'},u){\bf X}^{a+1}_{a,a+2}(u){\bf D}_{a+3}(u^{'})\nonumber\\ {\bf D}_{a+1}(u^{'}){\bf X}^{a+1}_{a,a+2}(u)&=&f_{1a}(u^{'},u) {\bf X}^{a+1}_{a,a+2}(u){\bf D}_{a+1}(u^{'}) +f_{2a}(u^{'},u){\bf X}^{a+1}_{a,a+2}(u^{'}){\bf D}_{a+1}(u)\nonumber\\ &+&f_{3a}(u^{'},u){\bf X}^{a+1}_{a,a+2}(u^{'}){\bf U}_{a+1}(u)\;, \label{eq:nuse1} \end{eqnarray} where \begin{eqnarray*} g_{1a}(u^{'},u)&=&-\frac{\sin(a\gamma)\sin((a+1)\gamma) \sin(\gamma-u^{'}+u)\sin(2\gamma-u^{'}-u)} {\sin((a-1)\gamma)\sin((a+2)\gamma)\sin(\gamma-u^{'}-u)\sin(u^{'}-u)}\\ g_{2a}(u^{'},u)&=&\frac{\sin\gamma\sin((a+1)\gamma) \sin(a\gamma+u^{'}-u)\sin(2\gamma-u^{'}-u)} {\sin((a-1)\gamma)\sin((a+2)\gamma)\sin(\gamma-u^{'}-u)\sin(u^{'}-u)}\\ g_{3a}(u^{'},u)&=&-\frac{\sin\gamma\sin((a+1)\gamma) \sin(a\gamma+u^{'}+u)\sin(2\gamma-u^{'}+u)} {\sin((a-1)\gamma)\sin((a+2)\gamma)\sin(\gamma-u^{'}-u)\sin(u^{'}-u)}\\ g_{4a}(u^{'},u)&=&\frac{\sin\gamma\sin(2\gamma) \sin(a\gamma+u^{'}+u)\sin((a+1)\gamma+u^{'}-u)} {\sin((a-1)\gamma)\sin((a+2)\gamma)\sin(\gamma-u^{'}-u)\sin(u^{'}-u)}\\ f_{1a}(u^{'},u)&=&-\frac{\sin(u^{'}+u) \sin(\gamma+u^{'}-u)} {\sin(\gamma-u^{'}-u)\sin(u^{'}-u)}\\ f_{2a}(u^{'},u)&=&\frac{\sin\gamma\sin(u^{'}+u) \sin((a+2)\gamma-u^{'}+u)} {\sin((a+2)\gamma)\sin(\gamma-u^{'}-u)\sin(u^{'}-u)}\\ f_{3a}(u^{'},u)&=&-\frac{\sin\gamma\sin((a+2)\gamma-u^{'}-u)} {\sin((a+2)\gamma)\sin(\gamma-u^{'}-u)}\\ \end{eqnarray*} The action of the transfer matrix on the state $\psi_a^b(\vec{\lambda})$ can be evaluated by commuting ${\bf U}_{b-1}(u)$ and ${\bf D}_{b+1}(u)$ through ${\bf X}^{a+N-i}_{a+N-i-1,a+N+i-1}(\lambda_i)$. However, the presence of the $g_{4a}(u^{'},u)$ term in the first commutation relation complicates matters considerably. Like the vertex case, it is desirable to define a new operator \begin{equation} \tilde{\bf U}_{a-1}(u)\equiv {\bf U}_{a-1}(u)+h_a(u){\bf D}_{a+1}(u) \end{equation} with \[h_a(u)=-\frac{\sin\gamma\sin((a-1)\gamma+2u)}{\sin((a-1)\gamma) \sin(\gamma-2u)}\;.\] so that the operators $\tilde{\bf U}_{a-1}(u),{\bf D}_{a+1}(u)$ satisfy simplified relations \begin{eqnarray} \tilde{\bf U}_{a-1}(u^{'}){\bf X}^{a+1}_{a,a+2}(u)&=&\alpha_{1a}(u^{'},u) {\bf X}^{a+1}_{a,a+2}(u)\tilde{\bf U}_{a+1}(u^{'}) +\alpha_{2a}(u^{'},u){\bf X}^{a+1}_{a,a+2}(u^{'}) \tilde{\bf U}_{a+1}(u)\nonumber\\ &&\mbox{}+\alpha_{3a}(u^{'},u){\bf X}^{a+1}_{a,a+2}(u^{'}){\bf D}_{a+3}(u)\\ {\bf D}_{a+1}(u^{'}){\bf X}^{a+1}_{a,a+2}(u)&=&\beta_{1a}(u^{'},u) {\bf X}^{a+1}_{a,a+2}(u){\bf D}_{a+1}(u^{'}) +\beta_{2a}(u^{'},u){\bf X}^{a+1}_{a,a+2}(u^{'}){\bf D}_{a+1}(u)\nonumber\\ &&\mbox{}+\beta_{3a}(u^{'},u){\bf X}^{a+1}_{a,a+2}(u^{'}) \tilde{\bf U}_{a+1}(u)\;, \end{eqnarray} where \begin{eqnarray*} \alpha_{1a}(u^{'},u)&=&-\frac{\sin(a\gamma)\sin((a+1)\gamma) \sin(\gamma-u^{'}+u)\sin(2\gamma-u^{'}-u)} {\sin((a-1)\gamma)\sin((a+2)\gamma)\sin(\gamma-u^{'}-u)\sin(u^{'}-u)}\\ \alpha_{2a}(u^{'},u)&=&\frac{\sin\gamma\sin(a\gamma) \sin((a+1)\gamma+u^{'}-u)\sin(2\gamma-2u)} {\sin((a-1)\gamma)\sin((a+2)\gamma)\sin(\gamma-2u^{'})\sin(u^{'}-u)}\\ \alpha_{3a}(u^{'},u)&=&-\frac{\sin\gamma\sin(a\gamma) \sin(2u)\sin(a\gamma+u^{'}+u)\sin(2\gamma-2u^{'})} {\sin((a-1)\gamma)\sin((a+1)\gamma)\sin(\gamma-u^{'}-u)\sin(\gamma-2u^{'})}\\ \beta_{1a}(u^{'},u)&=&-\frac{\sin(u^{'}+u) \sin(\gamma+u^{'}-u)} {\sin(\gamma-u^{'}-u)\sin(u^{'}-u)}\\ \beta_{2a}(u^{'},u)&=&\frac{\sin\gamma\sin(2u) \sin((a+1)\gamma-u^{'}+u)} {\sin((a+1)\gamma)\sin(\gamma-2u)\sin(u^{'}-u)}\\ \beta_{3a}(u^{'},u)&=&-\frac{\sin\gamma\sin((a+2)\gamma-u^{'}-u)} {\sin((a+2)\gamma)\sin(\gamma-u^{'}-u)}\;.\\ \end{eqnarray*} Moreover the eigenvalue $\tilde{\cal U}_a^N(u)$ of $\tilde{\bf U}_{a-1}(u)$ on $\omega_a^{a+N}$ is much simplified and given by \begin{equation} \tilde{\cal U}_a^N(u)=-\frac{\sin(a\gamma)\sin(2u) \sin(\xi_a+(a-1)\gamma+u)\sin(\xi_a+\gamma-u)} {\sin((a-1)\gamma)\sin(\gamma-2u)\sin(\xi_a+a\gamma+u) \sin(\xi_a-u)}\prod_{j=1}^N {\cal A}_{a+j}(u_j)U_{a-1}(u) \end{equation} where use has been made of the relation \begin{equation} {\cal B}_{a+i}(u_i)+h_{a+i}(u){\cal C}_{a+i}(u_i)=h_{a+i-1}(u) {\cal A}_{a+i}(u_i). \end{equation} Another important relation we need is \begin{equation} {\bf X}^{a+1}_{a,a+2}(u){\bf X}^{a+3}_{a+2,a+4}(u^{'})= {\bf X}^{a+1}_{a,a+2}(u^{'}){\bf X}^{a+3}_{a+2,a+4}(u) \end{equation} which implies that the Bethe ansatz state $\psi_a^b(\vec{\lambda})$ is a symmetric function in the $\lambda_i$'s, which is useful in obtaining the Bethe ansatz equation. It is now straight forward to compute the action of $\tilde{\bf U}_{b-1}(u),{\bf D}_{b+1}(u)$ on $\psi_a^b(\vec{\lambda})$ giving \begin{equation} \begin{array}{rl} \tilde{\bf U}_{b-1}(u)\psi_a^b(\vec{\lambda})&=\alpha_{1b}(u,\lambda_1) \alpha_{1b+2}(u,\lambda_2)\cdots\alpha_{1 a+N-2}(u,\lambda_{M}) \tilde{\cal U}_a^N(u)\psi_a^b(\vec{\lambda})\\ +&\left[\alpha_{2b}(u,\lambda_1)\alpha_{1b+2}(\lambda_1,\lambda_2) \cdots\alpha_{1 a+N-2}(\lambda_1,\lambda_{M}) \tilde{\cal U}_a^N(\lambda_1)\right.\\ +&\left.\alpha_{3b}(u,\lambda_1)\beta_{1b+2}(\lambda_1,\lambda_2) \cdots\beta_{1 a+N-2}(\lambda_1,\lambda_{M}) {\cal D}_a^N(\lambda_1)\right]\psi_a^b(\vec{\lambda};\lambda_1)\\ +&\cdots\\ +&\left[\alpha_{2b}(u,\lambda_M)\alpha_{1b+2}(\lambda_M,\lambda_2) \cdots\alpha_{1 a+N-2}(\lambda_M,\lambda_{M-1}) \tilde{\cal U}_a^N(\lambda_M)\right.\\ +&\left.\alpha_{3b}(u,\lambda_M)\beta_{1b+2}(\lambda_M,\lambda_2) \cdots\beta_{1 a+N-2}(\lambda_M,\lambda_{M-1}) {\cal D}_a^N(\lambda_M)\right]\psi_a^b(\vec{\lambda};\lambda_M)\\ \end{array}\end{equation} and \begin{equation} \begin{array}{rl} {\bf D}_{b+1}(u)\psi_a^b(\vec{\lambda})&=\beta_{1b}(u,\lambda_1) \beta_{1b+2}(u,\lambda_2)\cdots\beta_{1 a+N-2}(u,\lambda_{M}) {\cal D}_a^N(u)\psi_a^b(\vec{\lambda})\\ +&\left[\beta_{2b}(u,\lambda_1)\beta_{1b+2}(\lambda_1,\lambda_2) \cdots\beta_{1 a+N-2}(\lambda_1,\lambda_{M}) {\cal D}_a^N(\lambda_1)\right.\\ +&\left.\beta_{3b}(u,\lambda_1)\alpha_{1b+2}(\lambda_1,\lambda_2) \cdots\alpha_{1 a+N-2}(\lambda_1,\lambda_{M}) \tilde{\cal U}_a^N(\lambda_1)\right]\psi_a^b(\vec{\lambda};\lambda_1)\\ +&\cdots\\ +&\left[\beta_{2b}(u,\lambda_M)\beta_{1b+2}(\lambda_M,\lambda_2) \cdots\beta_{1 a+N-2}(\lambda_M,\lambda_{M-1}) {\cal D}_a^N(\lambda_M)\right.\\ +&\left.\beta_{3b}(u,\lambda_M)\alpha_{1b+2}(\lambda_M,\lambda_2) \cdots\alpha_{1 a+N-2}(\lambda_M,\lambda_{M-1}) \tilde{\cal U}_a^N(\lambda_M)\right]\psi_a^b(\vec{\lambda};\lambda_M)\;,\\ \end{array}\end{equation} where the state $\psi_a^b(\vec{\lambda};\lambda_i)$ is defined as in eqn.(\ref{eq:state}) with $\lambda_i$ replaced by $u$. Combining the above in the transfer matrix (see eqn.(\ref{eq:comtran})), \begin{eqnarray} {\bf T}(u)&\propto&\frac{\sin(2\gamma-2u) \sin(\xi_b-u)\sin(\xi_b+b\gamma+u)D_{b+1}(\gamma-u)} {\sin(\gamma-2u)\sin(\xi_b+\gamma-u)\sin(\xi_b+(b-1)\gamma+u)} {\bf D}_{b+1}(u)\nonumber\\ &&\mbox{}+\frac{\sin((b-1)\gamma) U_{b-1}(\gamma-u)}{\sin(b\gamma)} \tilde{\bf U}_{b-1}(u)\;, \end{eqnarray} the state $\psi_a^b(\vec{\lambda})$ is an eigenstate of the transfer matrix if the coefficients of $\psi_a^b(\vec{\lambda};\lambda_i),\;i=1,\cdots,M$ vanish, which gives the following Bethe ansatz equation to be satisfied for $\lambda_i$'s \begin{equation} \begin{array}{l} \displaystyle \frac{\textstyle\sin(\xi_b-\gamma+\lambda_i)\sin(\xi_b+(b+1)\gamma-\lambda_i) \sin(\xi_a+\gamma-\lambda_i)\sin(\xi_a+(a-1)\gamma+\lambda_i)} {\textstyle\sin(\xi_b-\lambda_i)\sin(\xi_b+b\gamma+\lambda_i) \sin(\xi_a+\lambda_i)\sin(\xi_a+a\gamma-\lambda_i)}\\ \\ = \displaystyle\prod_{\begin{array}{ccc} \scriptstyle j&\scriptstyle =&\scriptstyle 1\\\scriptstyle j&\scriptstyle \neq& \scriptstyle i\end{array}}^{M} \frac{\textstyle\sin(\lambda_i+\lambda_j)\sin(\gamma +\lambda_i-\lambda_j)}{\textstyle\sin(2\gamma-\lambda_i-\lambda_j) \sin(\gamma-\lambda_i+\lambda_j)}\prod_{k=1}^{N} \frac{\textstyle\sin(\gamma-\lambda_i+u_k) \sin(\gamma-\lambda_i-u_k)}{\textstyle\sin(\lambda_i-u_k) \sin(\lambda_i+u_k)}\;,\\ \hspace{8cm}\mbox{ for } i=1,\cdots,M \end{array}\label{eq:bae} \end{equation} and the eigenvalue $\Lambda_a^b(u,\vec{u};\vec{\lambda})$ of the transfer matrix is given by \begin{equation}\begin{array}{rcl} \displaystyle\Lambda_a^b(u,\vec{u};\vec{\lambda})&=&{\cal R}(\gamma-u){\cal R}_N(u) U_{b-1}(\gamma-u)D_{a+1}(u)\\ &&\displaystyle\times\left(\frac{\sin(2\gamma-2u)\sin(\xi_b-u) \sin(\xi_b+b\gamma+u)}{\sin(\gamma-2u)\sin(\xi_b-\gamma+u) \sin(\xi_b+(b+1)\gamma-u)}\right.\\ &&\displaystyle\times\prod_{i=1}^{M}\frac{\sin(u+\lambda_i) \sin(\gamma+u-\lambda_i)}{\sin(\lambda_i-u) \sin(\gamma-u-\lambda_i)}\prod_{k=1}^N\sin(\gamma-u+u_k)\sin(\gamma-u-u_k)\\ &-&\displaystyle\frac{\sin(2u)\sin(\xi_a+\gamma-u) \sin(\xi_a+(a-1)\gamma+u)}{\sin(\gamma-2u)\sin(\xi_a+u) \sin(\xi_a+a\gamma-u)}\\ &&\displaystyle\left.\times\prod_{i=1}^{M}\frac{\sin(u+\lambda_i-2\gamma) \sin(\gamma-u+\lambda_i)}{\sin(u-\lambda_i)\sin(\gamma-u-\lambda_i)} \prod_{k=1}^N\sin(u-u_k)\sin(u+u_k)\right)\;\label{eq:eigen} \end{array}\end{equation} In the above analysis, $\xi_a$ and $\xi_b$ are respectively the free parameters associated with the boundary $R$-matrices at the bottom and top boundaries. They need not be related at all, hence more generally one should write them as $\xi_{a}^{-}$ and $\xi_b^{+}$ to distinguish their origins. \section{Discussion and open problems} In this paper, we present the general trigonometric RSOS/SOS solution to the BYBE. By comparing them with the corresponding solution in the vertex representation, one may be able to obtain useful information about the vertex-SOS transformation matrix for the BYBE \cite{zou}. Indeed, for the diagonal solution, in the limit $\xi_a\rightarrow\pm i\infty$, $U_{a-1}$ and $D_{a+1}$ are equal so they contribute as an overall factor for the transfer matrix, since each bulk weight $S_{db}^{ac}$ is invariant under the action of the quantum group ${\rm U}_q sl(2)$ symmetry ($q=e^{i\gamma}$), the transfer matrix possesses the quantum symmetry. In this limit, the vertex-SOS transformation \cite{pas} is well known and has a precise meaning in terms of the Clebsch-Gordan coefficients of ${\rm U}_q sl(2)$. Thus for generic $\xi_a$, the vertex-SOS transformation can be considered as an extension of that of the ${\rm U}_q sl(2)$ case. The SOS/RSOS solutions indicate that the diagonal solution is not a special case of non-diagonal solution, in particular, there is no way of adjusting the free parameters to make all the off-diagonal scattering weights $X^a_{bc}$ vanish. The diagonal solution is the most favorable case to be studied as we have demonstrated how the transfer matrix of the SOS lattice built up from this solution may be diagonalized. On the other hand, it is not obvious how such method can be applied to the RSOS case. In the limit $\pi/\gamma$ becomes a rational number, solutions to the Bethe ansatz equation for the SOS model should contain those for the RSOS. However, except at the special point where there is a ${\rm U}_qsl(2)$ symmetry \cite{dev2}, it is not clear how the RSOS solutions may be extracted. It would be an interesting challenge to extend the idea to any $\xi_a$. The RSOS model is indeed a very interesting case to consider; it has been shown that for a different geometry where bulk faces are oriented at an angle $45^{\rm o}$ with respective to the boundaries, the RSOS boundary condition where all spins have height $a$ corresponds in the continuum limit to the boundary conformal state $|\tilde{h}_{1,a}\rangle$ and that with all boundary spins and their neighbors have respectively heights $a,a+1$ corresponds to $|\tilde{h}_{a,1}\rangle$ \cite{bau}. The lattice model that we considered has a different gemeotry from that in \cite{bau}, however, we believe that the difference is not significant in the scaling limit. Then the former boundary condition in fact can be obtained as the $\xi_a\rightarrow\pm i\infty$ limit of the RSOS diagonal solution since the weights $U_{a-1},D_{a+1}$ become the same. While the latter boundary condition can be obtained with $\xi_a=u$. So for generic $\xi_a$, the RSOS diagonal solution is in fact a mixture of the two above-mentioned boundary conditions and it would be interesting to examine which boundary conformal state it corresponds to in the scaling limit. Similarly, the non-diagonal solution, which gives to some extent a free boundary-like condition, may correspond to boundary conformal state $|\tilde{h}_{r,s}\rangle$ with $r,s\neq 1$ \cite{card} in the scaling limit. As a scattering theory, our results should be related to the conformal boundary conditions of the boundary minimal CFT and perturbations by relevant operators in the bulk and on the boundary. The non-diagonal solution, which has no free parameter, can be interpreted as the bulk $\Phi_{13}$ perturbation of the CFT with free boundary condition where all possible spins are allowed on the boundary. Additional integrable boundary perturbations introduce CDD-like factor in the scattering amplitudes. Each diagonal solution describes a perturbed CFT with a fixed boundary condition where the boundary ${\bf B}_a$ has fixed spin $a$. The parameter $\xi_a$ in the solution should be related to the coupling constant of the boundary perturbing field. It is an open problem to relate our solutions to the specific conformal boundary conditions and boundary perturbations of generic minimal CFTs. The case $p=4$ has been analyzed in \cite{chim}. One can also generalize our results to the coset CFTs ${\rm SU}(2)_k\otimes {\rm SU}(2)_l/{\rm SU}(2)_{k+l}$ perturbed by the least relevant operator in the bulk and by some boundary fields. The bulk-scattering matrices are given by \[ S=S_{{\rm RSOS}(k+2)}\otimes S_{{\rm RSOS}(l+2)}\;,\] where $S_{{\rm RSOS}(p)}$ is the RSOS $S$-matrix of the kinks. In this theory, particles carry two sets of RSOS spins and can be represented as $\vert K_{ab}\rangle\otimes\vert K_{a^{'}b^{'}}\rangle$ \cite{bl}. For the BYBE eqn.(\ref{eq:bybe}) with the above bulk scattering matrix, the boundary $R$-matrix given by \[ R=R_{{\rm RSOS}(k+2)}\otimes R_{{\rm RSOS}(l+2)}\;,\] is a solution, where $R_{{\rm RSOS}(p)}$ is the $R$-matrix given in eqn.(\ref{eq:boundary}). In particular, with $k=2$, this is the $N=1$ super CFTs with boundary perturbed by the least relevant operator and the $R$-matrix is given by that of the tri-critical Ising model tensored with that of the RSOS. To answer some of the questions raised, finding the Bethe ansatz equation and diagonalizing the transfer matrix for the RSOS case are the essential first step. It is quite likely that one needs an alternative method such as functional approach for this purpose. \section*{Acknowledgment} We thank A. LeClair who collaborated in the early stage of this work and H. Saleur for many useful discussion, WMK also acknowledges valuable discussion with P.Pearce. CA is supported in part by KOSEF 95-0701-04-01-3, 961-0201-006-2 and BSRI 95-2427 and WMK by a grant from KOSEF through CTP/SNU. \section*{Note added in proof} After completing this work, we learned that some of the results presented in section 2, 3 have also been independently obtained in \cite{pea,hou} as their trigonometric limit. \section*{Appendix} We present here the complete set of algebraic relations satisfied by the operators ${\bf X}^{a}_{bc}(u)$, ${\bf U}_a(u)$, ${\bf D}_a(u)$. These relations are obtained from the BYBE using eqn.(\ref{eq:operator}) for the boundary scattering matrix. Expanding the BYBE equation and considering the various allowed heights, we get \begin{equation} \begin{array}{l} c_{1a}(u_{+})c_{1a}(u_{-}){\bf U}_{a-1}{\bf U}_{a-1}^{'} +c_{1a}(u_{-}){\bf X}^{a-1}_{a-2,a}{\bf X}^{'a-1}_{a,a-2} +c_{3a}(u^{'},u){\bf U}_{a-1}{\bf D}_{a+1}^{'}\\ \mbox{}=c_{1a}(u_{+})c_{1a}(u_{-}){\bf U}_{a-1}^{'}{\bf U}_{a-1} +c_{1a}(u_{-}){\bf X}^{'a-1}_{a-2,a}{\bf X}^{a-1}_{a,a-2} +c_{3a}(u^{'},u){\bf D}_{a+1}^{'}{\bf U}_{a-1} \end{array} \end{equation} \begin{equation} \begin{array}{l} c_{2a}(u_{+})c_{2a}(u_{-}){\bf D}_{a+1}{\bf D}_{a+1}^{'} +c_{2a}(u_{-}){\bf X}^{a+1}_{a+2,a}{\bf X}^{'a+1}_{a,a+2} +c_{3a}(u^{'},u){\bf D}_{a+1}{\bf U}_{a-1}^{'}\\ \mbox{}=c_{2a}(u_{+})c_{2a}(u_{-}){\bf D}_{a+1}^{'}{\bf D}_{a+1} +c_{2a}(u_{-}){\bf X}^{'a+1}_{a,a+2}{\bf X}^{a+1}_{a+2,a} +c_{3a}(u^{'},u){\bf U}_{a-1}^{'}{\bf D}_{a+1} \end{array} \end{equation} \begin{equation} \begin{array}{l} c_{1a}f_{-}(u_{+}){\bf U}_{a-1}{\bf U}_{a-1}^{'} +c_{2a}f_{+}(u_{-}){\bf U}_{a-1}{\bf D}_{a+1}^{'} +f_{-}{\bf X}^{a-1}_{a,a-2}{\bf X}^{'a-1}_{a-2,a}\\ \mbox{}=c_{2a}f_{-}(u_{+}){\bf D}_{a+1}^{'}{\bf D}_{a+1} +c_{1a}f_{+}(u_{-}){\bf U}_{a-1}^{'}{\bf D}_{a+1} +f_{-}{\bf X}^{'a+1}_{a,a+2}{\bf X}^{a+1}_{a+2,a} \end{array} \end{equation} \begin{equation} \begin{array}{l} c_{2a}f_{-}(u_{+}){\bf D}_{a+1}{\bf D}_{a+1}^{'} +c_{1a}f_{+}(u_{-}){\bf D}_{a+1}{\bf U}_{a-1}^{'} +f_{-}{\bf X}^{a+1}_{a,a+2}{\bf X}^{'a+1}_{a+2,a}\\ \mbox{}=c_{1a}f_{-}(u_{+}){\bf U}_{a-1}^{'}{\bf U}_{a-1} +c_{2a}f_{+}(u_{-}){\bf D}_{a+1}^{'}{\bf U}_{a+1} +f_{-}{\bf X}^{'a-1}_{a,a-2}{\bf X}^{a-1}_{a-2,a} \end{array} \end{equation} \begin{equation} \begin{array}{l} c_{2a-2}(u_{+}){\bf X}^{a-1}_{a,a-2}{\bf X}^{'a-1}_{a-2,a} +{\bf U}_{a-1}{\bf U}_{a-1}^{'}=c_{2a-2}(u_{+}){\bf X}^{'a-1}_{a,a-2} {\bf X}^{a-1}_{a-2,a}\\ \mbox{ }+{\bf U}_{a-1}{\bf U}_{a-1}^{'} \end{array} \end{equation} \begin{equation} \begin{array}{l} c_{1a+2}(u_{+}){\bf X}^{a+1}_{a,a+2}{\bf X}^{'a+1}_{a+2,a} +{\bf D}_{a+1}{\bf D}_{a+1}^{'}= c_{1a+2}(u_{+}){\bf X}^{'a+1}_{a,a+2}{\bf X}^{a+1}_{a+2,a}\\ \mbox{ }+{\bf D}_{a+1}^{'}{\bf D}_{a+1} \end{array} \end{equation} \begin{equation} {\bf X}^{a}_{a-1,a+1}{\bf X}^{'a+2}_{a+1,a+3}= {\bf X}^{'a}_{a-1,a+1}{\bf X}^{a+2}_{a+1,a+3} \end{equation} \begin{equation} \begin{array}{l} {\bf X}^{a-1}_{a,a-2}{\bf D}_{a-1}^{'} +c_{1a}(u_{+}){\bf U}_{a-1}{\bf X}^{'a-1}_{a,a-2} =c_{1a}(u_{-}){\bf X}^{'a-1}_{a,a-2}{\bf D}_{a-1}\\ \mbox{}+c_{1a}(u_{+})c_{1a}(u_{-}){\bf U}_{a-1}^{'} {\bf X}^{a-1}_{a,a-2}+c_{3a}(u^{'},u){\bf D}_{a+1}^{'} {\bf X}^{a-1}_{a,a-2} \end{array} \end{equation} \begin{equation} \begin{array}{l} {\bf X}^{a+1}_{a,a+2}{\bf U}^{'}_{a+1} +c_{2a}(u_{+}){\bf D}_{a+1}{\bf X}^{'a+1}_{a,a+2} =c_{2a}(u_{-}){\bf X}^{'a+1}_{a,a+2}{\bf U}_{a+1}\\ \mbox{ }+c_{2a}(u_{+})c_{2a}(u_{-}){\bf D}^{'}_{a+1}{\bf X}^{a+1}_{a,a+2} +c_{3a}(u^{'},u){\bf U}_{a-1}^{'} {\bf X}^{a+1}_{a,a+2} \end{array}\label{eq:use1} \end{equation} \begin{equation} \begin{array}{l} c_{2a}(u_{-}){\bf U}_{a+1}{\bf X}^{'a+1}_{a+2,a} +c_{2a}(u_{+})c_{2a}(u_{-}){\bf X}^{a+1}_{a+2,a}{\bf D}^{'}_{a+1}\\ \mbox{ }+c_{3a}(u^{'},u){\bf X}^{a+1}_{a+2,a}{\bf U}_{a-1}^{'} ={\bf U}^{'}_{a+1}{\bf X}^{a+1}_{a+2,a}+c_{2a}(u_{+}){\bf X}^{'a+1}_{a+2,a} {\bf D}_{a+1} \end{array} \end{equation} \begin{equation} \begin{array}{l} c_{1a}(u_{-}){\bf D}_{a-1}{\bf X}^{'a-1}_{a,a-2} +c_{1a}(u_{+})c_{1a}(u_{-}){\bf X}^{a-1}_{a,a-2}{\bf U}^{'}_{a-1}\\ \mbox{ }+c_{3a}(u^{'},u){\bf X}^{a-1}_{a-2,a}{\bf D}_{a+1}^{'} ={\bf D}^{'}_{a-1}{\bf X}^{a-1}_{a-2,a}+c_{1a}(u_{+}) {\bf X}^{'a-1}_{a-2,a}{\bf U}_{a-1} \end{array} \end{equation} \begin{equation} \begin{array}{l} f_{+}{\bf U}_{a-1}{\bf X}^{'a+1}_{a,a+2}=f_{-}{\bf X}^{'a+1}_{a,a+2} {\bf U}_{a+1}+c_{1a}(u_{-})f_{+}{\bf U}_{a-1}^{'}{\bf X}^{a+1}_{a,a+2}\\ \mbox{ }+c_{2a}(u_{+})f_{-}{\bf D}^{'}_{a+1}{\bf X}^{a+1}_{a,a+2} \end{array} \end{equation} \begin{equation} \begin{array}{l} f_{+}{\bf D}_{a+1}{\bf X}^{'a-1}_{a,a-2}=f_{-}{\bf X}^{'a-1}_{a,a-2} {\bf D}_{a-1}+c_{2a}(u_{-})f_{+}{\bf D}_{a+1}^{'} {\bf X}^{a-1}_{a,a-2}\\ \mbox{ }+c_{1a}(u_{+})f_{-}{\bf U}^{'}_{a-1}{\bf X}^{a-1}_{a,a-2} \end{array} \end{equation} \begin{equation} \begin{array}{l} f_{-}{\bf U}_{a+1}{\bf X}^{'a+1}_{a+2,a}+c_{1a}(u_{-})f_{+} {\bf X}^{a+1}_{a+2,a}{\bf U}_{a-1}^{'}+ c_{2a}(u_{+})f_{-}{\bf X}^{a+1}_{a+2,a}{\bf D}^{'}_{a+1}\\ \mbox{ }=f_{+}{\bf X}^{'a+1}_{a+2,a}{\bf U}_{a-1} \end{array} \end{equation} \begin{equation} \begin{array}{l} f_{-}{\bf D}_{a-1}{\bf X}^{'a-1}_{a-2,a}+ c_{2a}(u_{-})f_{+}{\bf X}^{a-1}_{a-2,a}{\bf D}_{a+1}^{'} +c_{1a}(u_{+})f_{-}{\bf X}^{a-1}_{a-2,a}{\bf U}^{'}_{a-1}\\ \mbox{ }=f_{+}{\bf X}^{'a-1}_{a-2,a}{\bf D}_{a+1}\label{eq:use2} \end{array} \end{equation} where \[\begin{array}{lll} c_{1a}(u)&\equiv&\frac{\textstyle\sin\gamma\sin(a\gamma-u)} {\textstyle\sin(a\gamma)\sin(\gamma-u)}\\ c_{2a}(u)&\equiv&\frac{\textstyle\sin\gamma\sin(a\gamma+u)} {\textstyle\sin(a\gamma)\sin(\gamma-u)}\\ c_{3a}(u^{'},u)&\equiv&\frac{\textstyle\sin((a-1)\gamma)\sin((a+1)\gamma) \sin(u^{'}+u)\sin(u^{'}-u)} {\textstyle\sin^2(a\gamma)\sin(\gamma-u^{'}-u)\sin(\gamma-u^{'}+u)}\\ f_{\pm}&\equiv&\frac{\textstyle\sin(u^{'}\pm u)} {\textstyle\sin(\gamma-u^{'}\mp u)}\\ u_{\pm}&\equiv&u^{'}\pm u\;. \end{array}\] Here again, we abbreviate \[\begin{array}{ccc} {\bf U}_a&\equiv& {\bf U}_a(u)\\ {\bf U}_a^{'}&\equiv& {\bf U}_a(u^{'}) \end{array}\] and similarly for ${\bf D}_a$, ${\bf X}^a_{bc}$. Among them eqns.(\ref{eq:use1}),(\ref{eq:use2}) are of interest to us, with the help of the latter the former can be turned into the first equation in eqn.(\ref{eq:nuse1}) which is more convenient for the algebraic Bethe ansatz computation.
\section{INTRODUCTION} The investigation of semileptonic decays of $B$ mesons into light mesons is important for the determination of the Cabibbo-Kobayashi-Maskawa matrix element $V_{ub}$, which is the most poorly studied. At present the value of $V_{ub}$ is mainly determined from the endpoint of the lepton spectrum in semileptonic $B$-decays \cite{1}. Unfortunately, the theoretical interpretation of the endpoint region of the lepton spectrum in inclusive $B\to X_u\ell \bar \nu $ decays is very complicated and suffers from large uncertainties \cite{2}. The other way to determine $V_{ub}$ is to consider exclusive semileptonic decays $B\to \pi(\rho)e\nu$. These are the heavy-to-light transitions with a wide kinematic range. In contrast to the heavy-to-heavy transitions, here we can not expand matrix elements in the inverse powers of the final quark mass. It is also necessary to mention that the final meson has a large recoil momentum almost in the whole kinematical range. Thus the motion of final $\pi(\rho)$ meson should be treated relativistically. If we consider the point of maximum recoil of the final meson, we find that $\pi(\rho)$ bears the large relativistic recoil momentum $\vert{\bf \Delta}_{max}\vert$ of order $m_b/2$ and the energy of the same order. Thus at this kinematical point it is possible to expand the matrix element of the weak current both in inverse powers of $b$-quark mass of the initial $B$ meson and in inverse powers of the recoil momentum $\vert {\bf \Delta}_{max}\vert$ of the final $\pi(\rho)$ meson. As a result the expansion in powers $1/m_b$ arises for the $B\to \pi(\rho)$ semileptonic form factors at $q^2=0$, where $q^2$ is a momentum carried by the lepton pair. The aim of this paper is to realize such expansion in the framework of relativistic quark model. We show that this expansion considerably simplifies the analysis of exclusive $B\to \pi(\rho)e\nu$ semileptonic decays. Our relativistic quark model is based on the quasipotential approach in quantum field theory with the specific choice of the $q\bar q$ potential. It provides a consistent scheme for calculation of all relativistic corrections at a given order of $v^2/c^2$ and allows for the heavy quark $1/m_Q$ expansion. This model has been applied for the calculations of meson mass spectra \cite{3}, radiative decay widths \cite{4}, pseudoscalar decay constants \cite{5}, heavy-to-heavy semileptonic \cite{6} and nonleptonic \cite{7} decay rates. The heavy quark $1/m_Q$ expansion in our model for the heavy-to-heavy semileptonic transitions has been developed in \cite{8} up to $1/m_Q^2$ order. The results are in agreement with the model independent predictions of the heavy quark effective theory (HQET) \cite{9}. The $1/m_b$ expansion of rare radiative decay form factors of $B$ mesons has been carried out in \cite{10} along the same lines as in the present paper. We have briefly presented the results for $B\to\pi e\nu$ and $B\to \rho e\nu$ decays in ref.~\cite{10a}, where the expansion up to the first order in $1/m_b$ has been carried out. In the present paper we extend the analysis up to the second order and give a detailed discussion of the expansion method and results. The paper is organized as follows. The relativistic quark model is described in Sect.~2. In Sect.~3 we give the detailed description of the method of calculating decay matrix elements between heavy-heavy and heavy-light meson states, based on the quasipotential approach. We show that the heavy-to-heavy decay matrix elements can be expanded in inverse powers of the heavy quark masses at zero recoil of the final meson. On the other hand, the heavy-to-light decay matrix elements can be expanded in inverse powers of the initial heavy quark mass at the maximum recoil of the final light meson. These expansions permit the calculation of decay matrix elements with the account of relativistic effects. In Sect.~4 the method is applied to the calculation of the semileptonic $B\to\pi(\rho)e\nu$ decay form factors. Our numerical results for the form factors and decay rates are presented in Sect.~5. Therein we discuss the $q^2$-dependence of the form factors and the relations between semileptonic $B\to\rho e\nu$ and rare radiative $B\to\rho(K^*)\gamma$ decays. Sect.~6 contains our conclusions. The formulae for the form factors at the point of maximum recoil of the final light meson are given in Appendix. \section{RELATIVISTIC QUARK MODEL} In the quasipotential approach \cite{11} meson with the mass $M$ and relative momentum of quarks {\bf p} is described by a single-time quasipotential wave function $\Psi_M({\bf p})$, projected onto positive-energy states. This wave function satisfies the quasipotential equation \begin{equation} \label{1} \left(M-({\bf p}^2+m_a^2)^{1/2}-({\bf p}^2+m_b^2)^{1/2}\right) \Psi_M({\bf p}) = \int \frac{d^3 q}{(2\pi)^3} V({\bf p},{\bf q};M)\Psi_M({\bf q}), \end{equation} The quasipotential equation (\ref{1}) can be transformed into a local Schr\"odinger-like equation~\cite{12} \begin{equation} \label{2} \Big(\frac{b^2(M)}{2\mu_{R}}-\frac{{\bf p}^2}{2\mu_{R}}\Big)\Psi_{M}({\bf p})=\int\frac{d^3 q}{(2\pi)^3} V({\bf p,q};M)\Psi_{M}({\bf q}),\end{equation} where the relativistic reduced mass is \begin{equation} \label{3} \mu_{R}=\frac{M^4-(m^2_a-m^2_b)^2}{4M^3};\end{equation} and the square of the relative momentum on the mass shell is \begin{equation} \label{4} b^2(M)=\frac{[M^2-(m_a+m_b)^2][M^2-(m_a-m_b)^2]}{4M^2}, \end{equation} $m_{a,b}$ are the quark masses. While constructing the kernel of this equation $V({\bf p,q};M)$ --- the quasipotential of quark-antiquark interaction --- we have assumed that effective interaction is the sum of the one-gluon exchange term with the mixture of long-range vector and scalar linear confining potentials. We have also assumed that at large distances quarks acquire universal nonperturbative anomalous chromomagnetic moments and thus the vector long-range potential contains the Pauli interaction. The quasipotential is defined by \cite{3}: \begin{equation} \label{5} V({\bf p,q},M)=\bar{u}_a(p) \bar{u}_b(-p)\Big\{\frac{4}{3}\alpha_SD_{ \mu\nu}({\bf k})\gamma_a^{\mu}\gamma_b^{\nu}+V^V_{conf}({\bf k})\Gamma_a^{\mu} \Gamma_{b;\mu}+V^S_{conf}({\bf k})\Big\}u_a(q)u_b(-q),\end{equation} where $\alpha_S$ is the QCD coupling constant, $D_{\mu\nu}$ is the gluon propagator; $\gamma_{\mu}$ and $u(p)$ are the Dirac matrices and spinors; ${\bf k=p-q}$; the effective long-range vector vertex is \begin{equation} \label{6} \Gamma_{\mu}({\bf k})=\gamma_{\mu}+ \frac{i\kappa}{2m}\sigma_{\mu\nu}k^{\nu},\end{equation} $\kappa$ is the anomalous chromomagnetic quark moment. Vector and scalar confining potentials in the nonrelativistic limit reduce to \begin{equation} \label{7} V^V_{conf}(r)=(1-\varepsilon)(Ar+B),\ \ V^S_{conf}(r)=\varepsilon(Ar+B),\end{equation} reproducing $V_{nonrel}^{conf}(r)=V^S_{conf}+V^V_{conf}=Ar+B$, where $\varepsilon$ is the mixing coefficient. The explicit expression for the quasipotential with the account of the relativistic corrections of order $v^2/c^2$ can be found in ref. \cite{3}. All the parameters of our model: quark masses, parameters of linear confining potential $A$ and $B$, mixing coefficient $\varepsilon$ and anomalous chromomagnetic quark moment $\kappa$ were fixed from the analysis of meson masses \cite{3} and radiative decays \cite{4}. Quark masses: $m_b=4.88$ GeV; $m_c=1.55$ GeV; $m_s=0.50$ GeV; $m_{u,d}=0.33$ GeV and parameters of linear potential: $A=0.18$ GeV$^2$; $B=-0.30$ GeV have standard values for quark models. The value of the mixing coefficient of vector and scalar confining potentials $\varepsilon=-0.9$ has been primarily chosen from the consideration of meson radiative decays, which are very sensitive to the Lorentz-structure of the confining potential: the resulting leading relativistic corrections coming from vector and scalar potentials have opposite signs for the radiative Ml-decays \cite{4}. Universal anomalous chromomagnetic moment of quark $\kappa=-1$ has been fixed from the analysis of the fine splitting of heavy quarkonia ${ }^3P_J$- states \cite{3}. Recently we have considered the expansion of the matrix elements of weak heavy quark currents between pseudoscalar and vector meson states up to the second order in inverse powers of the heavy quark masses \cite{8}. It has been found that the general structure of leading, subleading and second order $1/m_Q$ corrections in our relativistic model is in accord with the predictions of HQET. The heavy quark symmetry and QCD impose rigid constraints on the parameters of the long-range potential of our model. The analysis of the first order corrections \cite{8} allowed to fix the value of effective long-range anomalous chromomagnetic moment of quarks $ \kappa =-1$, which coincides with the result, obtained from the mass spectra \cite{3}. The mixing parameter of vector and scalar confining potentials has been found from the comparison of the second order corrections to be $ \varepsilon =-1$. This value is very close to the previous one $\varepsilon =-0.9$ determined from radiative decays of mesons \cite{4}. Therefore, we have got QCD and heavy quark symmetry motivation for the choice of the main parameters of our model. The found values of $\varepsilon$ and $\kappa$ imply that confining quark-antiquark potential has predominantly Lorentz-vector structure, while the scalar potential is anticonfining and helps to reproduce the initial nonrelativistic potential. \section{MATRIX ELEMENTS OF ELECTROWEAK CURRENT BETWEEN HEAVY-HEAVY AND HEAVY-LIGHT MESON STATES} The matrix element of the local current $J$ between bound states in the quasipotential method has the form \cite{13} \begin{equation}\label{8} \langle M' \vert J_\mu (0) \vert M\rangle =\int \frac{d^3p\, d^3q}{(2\pi )^6} \bar \Psi_{M'}({\bf p})\Gamma _\mu ({\bf p},{\bf q})\Psi_M({\bf q}),\end{equation} where $M(M')$ is initial (final) meson, $\Gamma _\mu ({\bf p},{\bf q})$ is the two-particle vertex function and $\Psi_{M,M'}$ are the meson wave functions projected onto the positive energy states of quarks. This relation is valid for the general structure of the current $J=\bar Q'GQ$, where $G$ can be an arbitrary combination of Dirac matrices. The contributions to $\Gamma$ come from Figs.~1 and 2. Thus the vertex functions look like \begin{equation} \label{9}\Gamma^{(1)}({\bf p},{\bf q})=\bar u_{Q'}(p_1)G u_Q(q_1)(2\pi)^3\delta({\bf p}_2-{\bf q}_2),\end{equation} and \begin{eqnarray}\label{10} \Gamma^{(2)}({\bf p},{\bf q})&=&\bar u_{Q'}(p_1)\bar u_q(p_2) \Bigl\{G \frac{\Lambda_Q^{(-)}( k_1)}{\varepsilon_Q(k_1)+\varepsilon_Q(p_1)}\gamma_1^0V({\bf p}_2-{\bf q}_2)\nonumber \\ & &+V({\bf p}_2-{\bf q}_2)\frac{\Lambda_{Q'}^{(-)}(k_1')}{ \varepsilon_{Q'}(k_1')+ \varepsilon_{Q'}(q_1)}\gamma_1^0 G\Bigr\}u_Q(q_1) u_q(q_2),\end{eqnarray} where ${\bf k}_1={\bf p}_1-{\bf\Delta};\qquad {\bf k}_1'={\bf q}_1+{\bf\Delta};\qquad {\bf\Delta}={\bf p}_M-{\bf p}_{M'}; \qquad \varepsilon (p)=(m^2+{\bf p}^2)^{1/2}$; $$\Lambda^{(-)}(p)=\frac{\varepsilon(p)-\bigl( m\gamma ^0+\gamma^0({\bf \gamma p})\bigr)}{ 2\varepsilon (p)}.$$ and \begin{eqnarray*} p_{1,2}&=&\varepsilon_{1,2}(p)\frac{p_{M'}}{M'} \pm\sum_{i=1}^3 n^{(i)}(p_{M'})p^i,\\ q_{1,2}&=&\varepsilon_{1,2}(p)\frac{p_M}{M} \pm \sum_{i=1}^3 n^{(i)} (p_M)q^i,\end{eqnarray*} here $$ n^{(i)\mu}(p)=L_{p_i}^\mu=\left\{ \frac{p^i}{M},\ \delta_{ij}+ \frac{p^ip^j}{M(E+M)}\right\},$$ Note that the contribution $\Gamma^{(2)}$ is the consequence of the projection onto the positive-energy states. The form of the relativistic corrections resulting from the vertex function $\Gamma^{(2)}$ is explicitly dependent on the Lorentz-structure of $q\bar q$-interaction. The general structure of the current matrix element (\ref{8}) is rather complicated, because it is necessary to integrate both with respect to $d^3p$ and $d^3q$. The $\delta$-function in the expression (\ref{9}) for the vertex function $\Gamma^{(1)}$ permits to perform one of these integrations. As a result the contribution of $\Gamma^{(1)}$ to the current matrix element has usual structure and can be calculated without any expansion, if the wave functions of initial and final meson are known. The situation with the contribution $\Gamma^{(2)}$ is different. Here instead of $\delta$-function we have a complicated structure, containing the potential of $q\bar q$-interaction in meson. Thus in general case we cannot get rid of one of the integrations in the contribution of $\Gamma^{(2)}$ to the matrix element (\ref{8}). Therefore, it is necessary to use some additional considerations. The main idea is to expand the vertex function $\Gamma^{(2)}$, given by (\ref{10}), in such a way that it will be possible to use the quasipotential equation (\ref{1}) in order to perform one of the integrations in the current matrix element (\ref{8}). The realization of such expansion differs for the cases of heavy-to-heavy and heavy-to-light transitions. \subsection{Heavy-to-heavy decay matrix elements} At first we consider the heavy-to-heavy meson decays, such as semileptonic $B\to De\nu$ decays, and radiative transitions in quarkonia and $B^*\to B\gamma$, $D^*\to D\gamma$. Here we have two natural expansion parameters, which are the heavy quark masses in initial and final meson. The most convenient point for the expansion of vertex function $\Gamma^{(2)}$ in inverse powers of the heavy quark masses for semileptonic decays is the point of zero recoil of final meson, where ${\bf\Delta}=0$. For radiative decays the momentum transfer is fixed $\vert{\bf\Delta}\vert=\frac{M_M^2-M_{M'}^2}{2M}$. The difference of initial and final meson masses is proportional to the fine or hyperfine splitting and thus $\vert{\bf\Delta}\vert/M = o(1/M^2)$, so zero recoil is a good approximation. It is easy to see that $\Gamma^{(2)}$ contributes to the current matrix element at first order of $1/m_Q$ expansion for transitons between mesons consisting from heavy and light quarks ($B$, $D$ mesons) \cite{8} and at second order of $v/c$ expansion for mesons consisting from two heavy quarks of the same flavour (quarkonia $\Upsilon$, $J/\Psi$) \cite{4}. We limit our analysis to the consideration of the terms up to the second order in $1/m_Q$ or $v/c$ expansions. We substitute the Dirac matrices $G$ and spinors $u$ in the vertex function $\Gamma^{(2)}$ and consider the cases of Lorentz-scalar and Lorentz-vector (with Pauli term) $q\bar q$-interaction potential. Then we expand $\Gamma^{(2)}$ to the desired order and see that it is possible to integrate either with respect to $d^3p$ or $d^3q$ in the current matrix element (\ref{8}) using quasipotential equation (\ref{1}). Performing these integrations and taking the sum of the contributions of $\Gamma^{(1)}$ and $\Gamma^{(2)}$ we get the expression for the current matrix element, which contains the ordinary mean values between meson wave functions. Thus this matrix element can be easily calculated numerically if the meson wave functions are known. The described method has been applied to the calculations of heavy-to-heavy semileptonic decays in \cite{6,8} and radiative decays in \cite{4}. \subsection{Heavy-to-light decay matrix elements} Now we consider the heavy-to-light meson decays, such as semileptonic $B\to \pi(\rho)e\nu$ and rare radiative $B\to K^*\gamma$ decays. In these decays the final meson contains only light quarks ($u$, $d$, $s$), thus, in contrast to the heavy-to-heavy transitions, we cannot expand matrix elements in inverse powers of the final quark mass. The expansion of $\Gamma^{(2)}$ only in inverse powers of the initial heavy quark mass at ${\bf\Delta}=0$ does not allow to use the quasipotential equation for performing one of the integrations in corresponding current matrix element (\ref{8}). However, as it was already mentioned in the introduction, the final light meson has the large recoil momentum almost in the whole kinematical range. At the point of maximum recoil of final light meson \footnote{In the case of rare radiative decays the recoil momentum of final light meson is fixed at the maximum value ${\bf\Delta}_{max}$.} the large value of recoil momentum ${\bf\Delta}_{max}\sim m_Q/2$ allows for the expansion of decay matrix element in $1/m_Q$. The contributions to this expansion come both from the inverse powers of heavy $m_Q$ from initial meson and from inverse powers of the recoil momentum $|{\bf\Delta}_{max}|$ of the final light meson. The large value of recoil momentum $|{\bf\Delta}_{max}|$ permits to neglect ${\bf p}^2$ in comparison with ${\bf\Delta}_{max}^2$ in the light quark energy $\varepsilon_{q,Q'}(p+\Delta)$ in final meson in the expression for the matrix element originating from $\Gamma^{(2)}$. Such approximation corresponds to omitting terms of the third order in $1/m_Q$ expansion and is compatible with our analysis, which is carried out up to the second order. It is easy to see that we can now perform one of the integrations in the current matrix element (\ref{8}) using the quasipotential equation as in the case of heavy final meson. As a result we again get the expression for the current matrix element, which contains only the ordinary mean values between meson wave functions, but in this case at the point of maximum recoil of final light meson. This method has been applied to calculation of rare radiative decays of $B$ mesons in ref.~\cite{10} and in the next section we use it for consideration of $B\to\pi(\rho)e\nu$ semileptonic decays. \section{$B\to \pi(\rho)e\nu $ DECAY FORM FACTORS} \subsection{Decay form factors at $q^2=0$} The form factors of the semileptonic decays $B\to \pi e\nu$ and $B\to \rho e\nu$ are defined in the standard way as: \begin{eqnarray}\label{11}\langle\pi(p_\pi)\vert \bar q\gamma_\mu b\vert B(p_B) \rangle& =& f_+(q^2)(p_B+p_\pi)_\mu + f_-(q^2)(p_B-p_\pi)_\mu,\\ \label{12}\langle\rho(p_\rho,e)\vert \bar q\gamma_\mu(1-\gamma^5) b \vert B(p_B)\rangle &=& -(M_B+M_\rho)A_1(q^2)e^*_\mu + \frac{A_2(q^2)}{M_B+M_\rho} (e^* p_B)(p_B+p_\rho)_\mu\nonumber\qquad\qquad\hfil\hfill\\ & & +\frac{A_3(q^2)}{M_B+M_\rho}(e^* p_B)(p_B-p_\rho)_\mu + \frac{2V(q^2)}{M_B+M_\rho}i\epsilon_{\mu\nu \tau\sigma}{e^*}^\nu p_B^\tau p_\rho^\sigma,\end{eqnarray} where $q=p_B-p_{\pi(\rho)}$, $e$ is a polarization vector of $\rho $ meson. In the limit of vanishing lepton mass, the form factors $f_-$ and $A_3$ do not contribute to the decay rates and thus will not be considered. It is convenient to consider the decay $B\to\pi(\rho)e\nu$ in the $B$ meson rest frame. Then the wave function of the final $\pi(\rho)$ meson moving with the recoil momentum ${\bf\Delta}$ is connected with the wave function at rest by the transformation \cite{13} \begin{equation}\label{13}\Psi_{\pi(\rho)\,{\bf\Delta}}({\bf p})=D_q^{1/2}(R_{L{\bf\Delta}}^W)D_q^{1/2}(R_{L{ \bf\Delta}}^W)\Psi_{\pi(\rho)\,{\bf 0}}({\bf p}),\end{equation} where $D^{1/2}(R)$ is the well-known rotation matrix and $R^W$ is the Wigner rotation. The meson wave functions in the rest frame have been calculated by numerical solution of the quasipotential equation (\ref{2}) \cite{14}. However, it is more convenient to use analytical expressions for meson wave functions. The examination of numerical results for the ground state wave functions of mesons containing at least one light quark has shown that they can be well approximated by the Gaussian functions \begin{equation}\label{14}\Psi_M({\bf p})\equiv \Psi_{M\,{\bf 0}}({\bf p})=\Bigl({4\pi\over \beta_M^2} \Bigr)^{3/4}\exp\Bigl(-{{\bf p}^2\over 2\beta_M^2}\Bigr),\end{equation} with the deviation less than 5\%. The parameters are $$\beta_B=0.41\ {\rm GeV};\qquad\beta_{\pi(\rho)}=0.31\ {\rm GeV}.$$ Now we apply the method for calculation of decay matrix elements, described in the previous section. At the point of maximal recoil of final light meson \begin{equation}\label{15} |{\bf \Delta}_{max}| = \frac{M_B^2- M_{\pi(\rho)}^2}{2M_B}, \end{equation} we expand the vertex function $\Gamma^{(2)}$ for the Lorentz-scalar and Lorentz-vector (with Pauli term) $q\bar q$-interactions up to the second order in $1/m_b$. Then we substitute the vertex functions $\Gamma^{(1)}$ and $\Gamma^{(2)}$ in the matrix element (\ref{8}) and take into account the Lorentz transformation of the final meson wave function (\ref{13}). Performing one of the integrations in the current matrix element (\ref{8}) (using the $\delta$-function in $\Gamma^{(1)}$ and the quasipotential equation in the contribution of $\Gamma^{(2)}$) we get for the form factors of $B\to\pi e\nu$ and $B\to\rho e\nu$ decays the following expressions at $q^2=0$ point \begin{eqnarray} \label{16} f_+(0)&=&f_+^{(1)}(0)+\varepsilon f_{+}^{S(2)}(0)+(1-\varepsilon) f_{+}^{V(2)}(0),\\ \label{17} A_1(0)&=&A_1^{(1)}(0)+\varepsilon A_{1}^{S(2)}(0)+(1-\varepsilon) A_{1}^{V(2)}(0),\\ \label{18} A_2(0)&=&A_2^{(1)}(0)+\varepsilon A_{2}^{S(2)}(0)+(1-\varepsilon) A_{2}^{V(2)}(0),\\ \label{19} V(0)&=&V^{(1)}(0)+\varepsilon V^{S(2)}(0)+(1-\varepsilon) V^{V(2)}(0), \end{eqnarray} where $f_+^{(1)}$, $f_{+}^{S,V(2)}$, $A_{1,2}^{(1)}$, $A_{1,2}^{S,V(2)}$, $V^{(1)}$ and $V^{S,V(2)}$ are given in Appendix, the superscripts ``(1)" and ``(2)" correspond to Figs.~1 and 2, S and V --- to the scalar and vector potentials of $q\bar q$-interaction. \subsection{$1/m_b$ expansion for decay form factors} Let us proceed further and for the sake of consistensy carry out the complete expansion of form factors (\ref{16})--(\ref{19}) in inverse powers of $b$-quark mass. For this expansion we will use some model independent results obtained in HQET \cite{9}. In HQET the mass of $B$ meson has the following expansion in $1/m_b$ \cite{9} \begin{equation}\label{20}M_B=m_b+\bar\Lambda+\frac{\Delta m_B^2}{ 2m_b}+O\left(\frac{1}{ m_b^2}\right),\end{equation} where parameter $\bar\Lambda$ is the difference between the meson and quark masses in the limit of infinitely heavy quark mass. In our model $\bar\Lambda $ is equal to the mean value of light quark energy inside the heavy meson $\bar\Lambda=\langle\varepsilon_q\rangle_B\approx 0.54$~GeV \cite{8}. $\Delta m_B^2$ arises from the first-order power corrections to the HQET Lagrangian and has the form \cite{9}: \begin{equation}\label{21}\Delta m_B^2=-\lambda_1-3\lambda_2.\end{equation} The parameter $\lambda_1$ results from the mass shift due to the kinetic operator, while $\lambda_2$ parameterizes the chromomagnetic interaction \cite{9}. The value of spin-symmetry breaking parameter $\lambda_2$ is related to the vector-pseudoscalar mass splitting $$\lambda_2\approx {1\over 4}(M_{B^*}^2-M_B^2)=0.12\pm 0.01\ {\rm GeV}^2.$$ The parameter $\lambda_1$ is not directly connected with observable quantities. Theoretical predictions for it vary in a wide range: $\lambda_1=-0.30\pm 0.30 \ {\rm GeV}^2$ \cite{9,15}. In the limit $m_Q\to \infty$, meson wave functions become independent of the flavour of heavy quark. Thus the Gaussian parameter $\beta_B$ in (\ref{14}) should have the following expansion \cite{8} \begin{equation}\label{22}\beta_B=\beta-\frac{\Delta\beta^2}{ m_b}+O\left(\frac{1}{ m_b^2}\right),\qquad \beta \approx 0.42 \ {\rm GeV},\end{equation} where the second term breaks the flavour symmetry and in our model is equal to $\Delta\beta^2\approx 0.045\ {\rm GeV}^2$ \cite{8}. Substituting (\ref{20}) in (\ref{15}) and (\ref{a13}) we get the $1/m_b$ expansion of the recoil momentum and the energy of final vector meson: \begin{eqnarray}\label{23}\vert{\bf\Delta}_{max}\vert&=&\frac{m_b}{ 2}\left(1+\frac{1}{m_b}\bar\Lambda +\frac{1}{ m_b^2}\left(\frac{\Delta m_B^2}{ 2}-M_{\pi(\rho)}^2\right)\right)+O\left(\frac{1}{m_b^2}\right) ,\nonumber\\ E_{\pi(\rho)}&=&\frac{m_b}{ 2}\left(1+\frac{1}{ m_b}\bar \Lambda+\frac{1}{ m_b^2}\left(\frac{\Delta m_B^2}{2} +M_{\pi(\rho)}^2\right)\right)+O\left(\frac{1}{ m_b^2}\right).\end{eqnarray} Now we use the Gaussian approximation for the wave functions (\ref{14}). Then shifting the integration variable ${\bf p}$ in (\ref{a1})--(\ref{a12}) by $-\frac{\varepsilon_q}{ E_{\pi(\rho)}+M_{\pi(\rho})}{\bf\Delta}_{max}$, we can factor out the ${\bf\Delta}_{max}$ dependence of the meson wave function overlap in form factors $f_+$, $A_{1,2}$, $V$. The result can be written in the form \begin{equation}\label{24}f_+(0)={\cal F}_+({\bf\Delta}_{max}^2)\exp(-\zeta {\bf \Delta}_{max}^2),\end{equation} \begin{equation}\label{25}A_{1,2}(0)={\cal A}_{1,2}({\bf\Delta}_{max}^2)\exp(-\zeta {\bf \Delta}_{max}^2),\end{equation} \begin{equation}\label{26}V(0)={\cal V}({\bf\Delta}_{max}^2)\exp(-\zeta {\bf \Delta}_{max}^2),\end{equation} where $\vert{\bf\Delta}_{max}\vert$ is given by (\ref{15}) and \begin{equation}\label{27}\zeta{\bf\Delta}_{max}^2= \frac{2\tilde\Lambda^2 {\bf\Delta}_{max}^2}{(\beta_B^2+\beta_{\pi(\rho)}^2)(E_{\pi(\rho)}+ M_{\pi(\rho)})^2} =\frac{\tilde\Lambda^2}{ \beta_B^2 }\eta\left(\frac{M_B-M_{\pi(\rho)}}{M_B+M_{\pi(\rho)}}\right)^2, \end{equation} here $\eta = \frac{2\beta_B^2}{\beta_B^2+\beta_{\pi(\rho)}^2}$ and $\tilde\Lambda$ is equal to the mean value of light quark energy between $B$ and $\pi(\rho)$ meson states: \begin{equation}\label{28}\tilde\Lambda =\langle\varepsilon_q\rangle \approx 0.53\ {\rm GeV}, \end{equation} Expanding (\ref{27}) in powers of $1/m_b$ we get \begin{equation}\label{29}\zeta{\bf\Delta}_{max}^2 =\frac{\tilde\Lambda^2}{ \beta^2}\eta\left(1-4{M_{\pi(\rho)}\over m_b}\right)+O\left(\frac{1}{ m_b^2}\right). \end{equation} We see that the first term in this expansion for the decay into $\rho$ meson is large. Really, $4{M_\rho / m_b}\approx 0.63$. The value of this correction is also increased by the exponentiating in (\ref{24})--(\ref{26}). Therefore, we conclude that the first order correction in $1/m_b$ expansion for the form factors of $B\to\pi(\rho)e\nu$ decay, arising from the meson wave function overlap, is large.\footnote{The same situation occurs for the rare radiative $B$ decays \cite{10}.} Thus, taking into account that our method of calculating decay matrix elements does not require the expansion of the meson wave function overlap, we use unexpanded expression (\ref{27}) in the exponential of the form factors (\ref{24})--(\ref{26}). In contrast to the meson wave function overlap the factors ${\cal F}_1({\bf \Delta}_{max}^2) $, ${\cal A}_{1,2}({\bf\Delta}_{max}^2)$ and ${\cal V}({\bf\Delta}_{max}^2)$, which originate from the vertex functions $\Gamma^{(1),(2)}$ and Lorentz-transformation (\ref{13}) of the final meson wave function, have a well defined $1/m_b$ expansion. The first and second order corrections are small. Substituting the Gaussian wave functions (\ref{13}) in the expressions for the form factor (\ref{16})--(\ref{19}) and (\ref{a1})--(\ref{a12}), with the value of anomalous chromomagnetic quark moment $\kappa=-1$, and using (\ref{24})--(\ref{26}) and the expansions (\ref{20})--(\ref{23}), we get up to the second order in $1/m_b$ expansion: \nopagebreak \\ \medskip \nopagebreak \noindent a) $B\to\pi e\nu$ decay \nopagebreak \begin{eqnarray}\label{30}{\cal F}_+({\bf\Delta}_{max}^2)&=&{\cal F}_+^{(1)}({\bf\Delta}_{max}^2)+ \varepsilon{\cal F}_+^{S(2)}({\bf\Delta}_{max}^2)+(1-\varepsilon){\cal F}_+^{V(2)}({\bf \Delta}_{max}^2);\\ \label{31} {\cal F}_1^{(1)}({\bf\Delta}_{max}^2)&=&N\biggl\{1+\frac{1}{ m_b}X_1+\frac{1}{m_b^2}\biggl(Y_+-2\langle{\bf p}^2\rangle -\frac{3}{4}\tilde\Lambda^2\eta^2+\tilde\Lambda\eta(2m_q+M_\pi) \nonumber\\ & & +\frac{2}{3}\left\langle\frac{{\bf p}^2}{\bar\varepsilon_q+m_q} \right\rangle \left(2\tilde\Lambda\eta-m_q-2M_\pi- \frac{1}{2}\bar\Lambda\right) \biggr)\biggr\};\\ \label{32} {\cal F}_+^{S(2)}({\bf\Delta}_{max}^2) &=&N\frac{1}{m_b^2}\biggl\{ -2m_q(M_\pi-2\langle\bar\varepsilon_q\rangle) -\frac{1}{3}\left\langle \frac{{\bf p}^2}{\bar\varepsilon_q+m_q}\right\rangle(M_\pi(1-R) +2m_q(1-2R)-2\bar\Lambda R)\nonumber\\ & &+\frac{2}{3}\langle{\bf p}^2\rangle(1-2R)+\frac{1}{2}\tilde\Lambda \eta\left(2\langle\bar\varepsilon_q\rangle(1-2R)-M_\pi(1-R) +2\bar\Lambda R -\frac{1}{3}\left\langle\frac{{\bf p}^2}{\bar \varepsilon_q}\right\rangle\right)\biggr\};\\ \label{33} {\cal F}_+^{V(2)}({\bf\Delta}_{max}^2) &=&N \biggl\{-\frac{1}{m_b} (2RZ_1+(1+2R)Z_2) +\frac{1}{m_b^2}\biggl( -\frac{2}{3}m_qRZ_3 +4\langle{\bf p}^2\rangle\left(1-\frac{3}{4}R\right)\nonumber\\ & & -\left\langle\frac{{\bf p}^2}{\bar\varepsilon_q+m_q}\right\rangle \left( 2M_\pi \left(1-\frac{1}{3}R\right) +4m_q\left(1-\frac{5}{4}R\right) +\frac{5}{3}\bar\Lambda R\right)\nonumber\\ & & +\left(\left(2M_\pi +\eta\frac{\beta_\pi^2}{\beta^2}\frac{\Delta\beta^2}{\beta}\right) (1+2R)-2m_q(1-2R)\right)Z_2\biggr)\biggr\}; \end{eqnarray} \noindent b) $B\to\rho e\nu$ decay \nopagebreak \begin{eqnarray}\label{34}{\cal A}_1({\bf\Delta}_{max}^2)&=&{\cal A}_1^{(1)}({\bf\Delta}_{max}^2)+ \varepsilon{\cal A}_1^{S(2)}({\bf\Delta}_{max}^2)+(1-\varepsilon){\cal A}_1^{V(2)}({\bf \Delta}_{max}^2);\\ \label{35} {\cal A}_1^{(1)}({\bf\Delta}_{max}^2)&=&N\biggl\{1+\frac{1}{ m_b}X_-+\frac{1}{m_b^2}\biggl(Y_-+M_\rho(M_\rho-m_q) +\bar\Lambda(M_\rho+m_q)-\frac{1}{2}\tilde\Lambda\eta(5m_q+3M_\rho) \nonumber\\ & & +\frac{1}{3}\left\langle\frac{{\bf p}^2}{\bar\varepsilon_q+m_q} \right\rangle \left(2\tilde\Lambda\eta-m_q-3M_\rho\right) \biggr)\biggr\};\\ \label{36} {\cal A}_1^{S(2)}({\bf\Delta}_{max}^2) &=&N\biggl\{ \frac{2}{m_b}(M_\rho-2\langle\bar\varepsilon_q\rangle)+ \frac{1}{m_b^2}\biggl( -2(M_\rho+\bar\Lambda+3m_q)(M_\rho-2\langle\bar\varepsilon_q\rangle) \nonumber\\ & & -\frac{2}{3}\left\langle \frac{{\bf p}^2}{\bar\varepsilon_q+m_q}\right\rangle(M_\rho R +m_q(1+3R)+ \bar\Lambda(1+R)) +\frac{2}{3}\langle{\bf p}^2\rangle(1+3R)\nonumber\\ & &-\tilde\Lambda \eta\left( M_\rho(2+R) -\langle\bar\varepsilon_q\rangle(7+3R) +\bar\Lambda(3+ R) +\frac{1}{3}\left\langle\frac{{\bf p}^2}{\bar \varepsilon_q}\right\rangle(1+R)\right)\biggr\};\\ \label{37} {\cal A}_1^{V(2)}({\bf\Delta}_{max}^2) &=&N \biggl\{\frac{1}{m_b} (4RZ_1+2RZ_2) +\frac{1}{m_b^2}\biggl( \frac{4}{3}(m_q-3M_\rho)RZ_3 -10\langle{\bf p}^2\rangle\left(1-\frac{2}{3}R\right)\nonumber\\ & & +\left\langle\frac{{\bf p}^2}{\bar\varepsilon_q+m_q}\right\rangle \left( 2M_\rho \left(2+\frac{13}{3}R\right) +8m_q\left(1-\frac{4}{3}R\right) \right)\nonumber\\ & & +2\left(2m_q(1-R)-2M_\rho R +\eta\frac{\beta_\rho^2}{\beta^2}\frac{\Delta\beta^2}{\beta}R\right)Z_. \nonumber\\ & &+2\tilde\Lambda\eta\biggl(\left\langle\frac{1}{\bar\varepsilon_q +m_q}\right\rangle\tilde\Lambda\eta\left(M_\rho\left(1- \frac{5}{3}R\right) -\bar\Lambda +m_q\left(1-\frac{10}{3}R\right)\right)\nonumber\\ & &+2(3\bar\Lambda -M_\rho-\langle\bar\varepsilon_q\rangle) -\tilde\Lambda\eta\left(1 -\frac{10}{3}R\right) -\frac{10}{3}\left\langle\frac{{\bf p}^2} {\bar\varepsilon_q}\right\rangle\biggr)\biggr)\biggr\}; \end{eqnarray} \begin{eqnarray}\label{38}{\cal A}_2({\bf\Delta}_{max}^2)&=&{\cal A}_2^{(1)}({\bf\Delta}_{max}^2)+ \varepsilon{\cal A}_2^{S(2)}({\bf\Delta}_{max}^2)+(1-\varepsilon){\cal A}_2^{V(2)}({\bf \Delta}_{max}^2);\\ \label{39} {\cal A}_2^{(1)}({\bf\Delta}_{max}^2)&=&N\biggl\{1+\frac{1}{ m_b}X_-+\frac{1}{m_b^2}\biggl(Y_--M_\rho(3M_\rho-m_q) +\bar\Lambda(M_\rho-m_q)-\frac{1}{2}\tilde\Lambda\eta(M_\rho-7m_q) \nonumber\\ & & -2\tilde\Lambda^2\eta^2+\frac{1}{3}\left\langle\frac{{\bf p}^2}{\bar\varepsilon_q+m_q} \right\rangle \left(2\tilde\Lambda\eta-m_q+5M_\rho\right) \biggr)\biggr\};\\ \label{40} {\cal A}_1^{S(2)}({\bf\Delta}_{max}^2) &=& {\cal A}_1^{S(2)}({\bf\Delta}_{max}) +N\biggl\{ \frac{1}{m_b^2}4M_\rho(M_\rho-2\langle\bar\varepsilon_q \rangle)\biggr\};\\ \label{41} {\cal A}_2^{V(2)}({\bf\Delta}_{max}^2) &=&N \biggl\{\frac{1}{m_b} (4RZ_1+2RZ_2) +\frac{1}{m_b^2}\biggl( \frac{4}{3}(M_\rho-m_q)RZ_3 -2\langle{\bf p}^2\rangle\left(1-\frac{10}{3}R\right)\nonumber\\ & & +\left\langle\frac{{\bf p}^2}{\bar\varepsilon_q+m_q}\right\rangle \left(\frac{2}{3}M_\rho \left(1-11R\right) +\frac{2}{3}\bar\Lambda +2m_q(1+R) \right)\nonumber\\ & & +2\left(m_q(1+3R)-3M_\rho R +\eta\frac{\beta_\rho^2}{\beta^2}\frac{\Delta\beta^2}{\beta}R\right)Z_. \nonumber\\ & &+2\tilde\Lambda\eta\biggl(-\left\langle\frac{1}{\bar\varepsilon_q +m_q}\right\rangle\tilde\Lambda\eta\left(M_\rho\left(1+ \frac{5}{3}R\right) +\bar\Lambda +m_q\left(3+\frac{10}{3}R\right)\right)\nonumber\\ & &+3\bar\Lambda -M_\rho-\langle\bar\varepsilon_q\rangle +\tilde\Lambda\eta\left(3 +\frac{10}{3}R\right) -\frac{1}{3}\left\langle\frac{{\bf p}^2} {\bar\varepsilon_q}\right\rangle\biggr)\biggr)\biggr\}; \end{eqnarray} \begin{eqnarray}\label{42}{\cal V}({\bf\Delta}_{max}^2)&=&{\cal V}^{(1)}({\bf\Delta}_{max}^2)+ \varepsilon{\cal V}^{S(2)}({\bf\Delta}_{max}^2)+(1-\varepsilon){\cal V}^{V(2)}({\bf \Delta}_{max}^2);\\ \label{43} {\cal V}^{(1)}({\bf\Delta}_{max}^2)&=&N\biggl\{1+\frac{1}{ m_b}X_++\frac{1}{m_b^2}\biggl(Y_-+M_\rho(M_\rho-m_q) +\frac{1}{2}m_q^2 +\bar\Lambda(2m_q-M_\rho)+\frac{7}{2}\tilde\Lambda\eta M_\rho \nonumber\\ & &-4\frac{1}{3}\langle{\bf p}^2\rangle -\frac{1}{3}\left\langle\frac{{\bf p}^2}{\bar\varepsilon_q+m_q} \right\rangle (M_\rho-m_q) \biggr)\biggr\};\\ \label{44} {\cal V}^{S(2)}({\bf\Delta}_{max}^2) &=&N\biggl\{ -\frac{2}{m_b}(M_\rho-2\langle\bar\varepsilon_q\rangle)+ \frac{1}{m_b^2}\biggl( -2(M_\rho-\bar\Lambda-m_q)(M_\rho-2\langle\bar\varepsilon_q\rangle) \nonumber\\ & & +\frac{2}{3}\left\langle \frac{{\bf p}^2}{\bar\varepsilon_q+m_q}\right\rangle(M_\rho(1-R) +m_q(2-3R)- \bar\Lambda R) +\frac{2}{3}\langle{\bf p}^2\rangle(2-3R)\nonumber\\ & &+\tilde\Lambda \eta\left(- M_\rho(1+2R) +6\langle\bar\varepsilon_q\rangle R+ 2\bar\Lambda(1- R) -\frac{2}{3}\left\langle\frac{{\bf p}^2}{\bar \varepsilon_q}\right\rangle(2-R)\right)\biggr\};\\ \label{45} {\cal V}^{V(2)}({\bf\Delta}_{max}^2) &=&N \biggl\{\frac{1}{m_b} (8RZ_1-2(1-R)Z_2) +\frac{1}{m_b^2}\biggl(- \frac{8}{3}(M_\rho-m_q)RZ_3 +20\langle{\bf p}^2\rangle\nonumber\\ & & -\left\langle\frac{{\bf p}^2}{\bar\varepsilon_q+m_q}\right\rangle \left( 2M_\rho \left(\frac{11}{3}-4R\right) +\frac{22}{3}\bar\Lambda +4m_q\left(5+2R\right) \right)\nonumber\\ & & +2\left(\bar\Lambda +\left(M_\rho R +\eta\frac{\beta_\rho^2}{\beta^2}\frac{\Delta\beta^2}{\beta}\right) (1-R)+m_q(1+R)\right)Z_2 \nonumber\\ & &+\tilde\Lambda^2\eta^2\biggl(17- \left\langle\frac{1}{\bar\varepsilon_q +m_q}\right\rangle\left(\frac{16}{3}M_\rho +\frac{19}{3}\bar\Lambda +17m_q \right) \biggr)\biggr)\bigg\}, \end{eqnarray} where $N=\left(\frac{2\beta_B\beta_{\pi(\rho)}}{ \beta_B^2+\beta_{\pi(\rho)}^2}\right)^{3/2} =\left(\frac{\beta_{\pi(\rho)}}{\beta_B}\eta\right)^{3/2}$ is due to the normalization of Gaussian wave functions in (\ref{14}); $\bar\varepsilon_q =\sqrt{{\bf p}^2+m_q^2+\tilde\Lambda^2\eta^2} $, i.~e. the energies of light quarks in final light meson acquire additional contribution from the recoil momentum. The averaging $\langle\dots\rangle$ is taken over the Gaussian wave functions of $B$ and $\pi(\rho)$ mesons, so it can be carried out analytically. For example, \begin{equation}\label{q}\langle\bar\varepsilon_q\rangle= N^{-1}\int \frac{d^3p}{(2\pi)^3}\bar\Psi_{\pi(\rho)}({\bf p})\bar\varepsilon_q(p)\Psi_B({\bf p}) = \frac{1}{ \sqrt\pi}\frac{\bar m_q^2}{\beta_{\pi(\rho)}\sqrt\eta}e^zK_1(z), \end{equation} where $\bar m_q^2=m_q^2+\tilde\Lambda^2\eta^2$ and $K_1(z)$ is the modified Bessel function; $z={\bar m_q^2/(2\eta\beta_{\pi(\rho)}^2)}$. Analogous expressions can be obtained for the other matrix elements in (\ref{30})--(\ref{45}). We have introduced the following notations: \begin{eqnarray*} X_1&=&\frac{2}{3}\left\langle\frac{{\bf p}^2}{\bar\varepsilon_q +m_q}\right\rangle -\frac{1}{2}\tilde\Lambda\eta,\\ X_\pm&=&\frac{1}{3}\left\langle\frac{{\bf p}^2}{\bar\varepsilon_q +m_q}\right\rangle +\frac{1}{2}\tilde\Lambda\eta \pm(M_\rho-m_q),\\ Y_\pm &=& -\frac{11}{24}\langle{\bf p}^2\rangle +\frac{1}{2}(M_{\pi(\rho)}^2-m_q^2) \pm\frac{1}{2}\tilde\Lambda \eta^2\left(\frac{1}{4}\tilde\Lambda+ \frac{\beta_{\pi(\rho)}^2} {\beta^2}\frac{\Delta\beta^2}{\beta}\right),\\ Z_1& =& \frac{1}{3}\left\langle\frac{{\bf p}^2}{(\bar\varepsilon_q +m_q)^2}\right\rangle(\bar\Lambda +M_{\pi(\rho)}+3m_q) -\left\langle\frac{{\bf p}^2}{\bar\varepsilon_q +m_q}\right\rangle,\\ Z_2 &=& \tilde\Lambda\eta\left(\frac{1}{3}\left\langle \frac{{\bf p}^2}{\bar\varepsilon_q(\bar\varepsilon_q+m_q)}\right\rangle +\left\langle\frac{1}{\bar\varepsilon_q+m_q)}\right\rangle (\bar\Lambda+M_{\pi(\rho)}+3m_q)-3\right),\\ Z_3 &=& \left\langle\frac{{\bf p}^2}{(\bar\varepsilon_q +m_q)^2} \right\rangle(\bar\Lambda+M_{\pi(\rho)}+3m_q) \end{eqnarray*} and $$R=\frac{m_b}{\varepsilon_b(\Delta_{max})+m_b}=\frac{1}{\sqrt{5}+2}.$$ \section{RESULTS AND DISCUSSION} Using the parameters of Gaussian wave functions (\ref{14}) and the value of the mixing coefficient of vector and scalar confining potentials $\varepsilon=-1$ \cite{8} in the expressions (\ref{24}), (\ref{30})--(\ref{33}) for the $B\to \pi$ transition form factor $f_+(0)$ and the eqs.~(\ref{25}), (\ref{26}), (\ref{34})--(\ref{45}) for the $B\to \rho$ transition form factors $A_1(0)$, $A_2(0)$ and $V(0)$ we get \begin{eqnarray}\label{46} f_+^{B\to \pi}(0)&=&0.20\pm 0.02\qquad V^{B\to \rho}(0) =0.29\pm 0.03\nonumber\\ A_1^{B\to\rho}(0)&=&0.26\pm 0.03\qquad A_2^{B\to \rho}(0)=0.31\pm 0.03.\end{eqnarray} The theoretical uncertainty in (\ref{46}) results mostly from the approximation of the wave functions by Gaussians (\ref{14}) and does not exceed 10\% of form factor values. In ref.~\cite{10a} we have presented the results for $B\to\pi(\rho)e\nu$ decay form factors up to the first order in $1/m_b$ expansion. The found values of form factors \cite{10a} are very close to (\ref{46}), this indicates that the second order correctons in $1/m_Q$ are small (less than 5\% of form factor values). We compare our results (\ref{46}) for the form factors of $B\to \pi(\rho)e\nu$ decays with the predictions of quark models \cite{16,17}, QCD sum rules \cite{18,19,19a} and lattice calculations \cite{l1,l2} in Table 1. There is an agreement between our value of $f_+^{B\to \pi}(0)$ and QCD sum rule and lattice predictions. Our $B\to \rho e\nu$ form factors agree with lattice and QCD sum rule ones \cite{19a}, while they are approximately 1.5 times less than QCD sum rule results of refs.~\cite{18,19}. To calculate the $B\to \pi(\rho)$ semileptonic decay rates it is necessary to determine the $q^2$-dependence of the form factors. Analysing the ${\bf\Delta}_{max}^2$ dependence of the expressions (\ref{a1})--(\ref{a12}), (\ref{24})--(\ref{27}) for the form factors $f_+$, $A_1$, $A_2$ and $V$, we find that the $q^2$-dependence of these form factors near $q^2=0$ is given by \begin{eqnarray} \label{47} f_+(q^2)&=&\frac{M_B+M_\pi}{2\sqrt{M_B M_\pi}}\tilde\xi(w){\cal F}_+({\bf\Delta}_{max}^2),\\ \label{48} A_1(q^2)&=&\frac{2\sqrt{M_B M_\rho}}{M_B+M_\rho}\frac{1}{2}(1+w) \tilde\xi(w){\cal A}_1 ({\bf\Delta}_{max}^2),\\ \label{49} A_2(q^2)&=&\frac{M_B+M_\rho}{2\sqrt{M_B M_\rho}}\tilde\xi(w){\cal A}_2 ({\bf\Delta}_{max}^2),\\ \label{50} V(q^2)&=&\frac{M_B+M_\rho}{2\sqrt{M_B M_\rho}}\tilde\xi(w){\cal V} ({\bf\Delta}_{max}^2),\end{eqnarray} where $w=\frac{M_B^2+M_{\pi(\rho)}^2-q^2}{2M_B M_{\pi(\rho)}}$; ${\cal F}_+({\bf\Delta}_{max}^2)$ and ${\cal A}_{1,2}({\bf \Delta}_{max}^2)$, ${\cal V}({\bf\Delta}_{max}^2)$ are defined by (\ref{24})--(\ref{26}), (\ref{30})--(\ref{45}). We have introduced the function \begin{equation} \label{51} \tilde\xi(w)=\left({2\over w+1}\right)^{1/2}\exp \left(-\eta\frac{\tilde\Lambda^2}{\beta^2_B}\frac{w-1}{w+1}\right), \end{equation} which in the limit of infinitely heavy quarks in the initial and final mesons coinsides with the Isgur-Wise function of our model \cite{8}. In this limit eqs.~(\ref{47})--(\ref{50}) reproduce the leading order prediction of HQET~\cite{9}. It is important to note that the form factor $A_1$ in (\ref{48}) has a different $q^2$-dependence than the other form factors (\ref{47}), (\ref{49}), (\ref{50}). In the quark models it is usually assumed the pole \cite{16} or exponential \cite{17} $q^2$-behaviour for all form factors. However, the recent QCD sum rule analysis indicates that the form factor $A_1$ has $q^2$-dependence different from other form factors \cite{18,19,19a}. In \cite{19} it even decreases with the increasing $q^2$ as \begin{equation} \label{52} A_1(q^2)\simeq \left(1-\frac{q^2}{M_b^2}\right)A_1(0) \simeq \frac{2M_B M_\rho}{(M_B+M_\rho)^2}(1+w)A_1(0). \end{equation} Such behaviour corresponds to replacing $\tilde\xi(w)$ in (\ref{48}) by $\tilde\xi(w_{max})$. We have calculated the decay rates of $B\to \pi(\rho)e\nu$ using our form factor values at $q^2=0$ and the $q^2$-dependence (\ref{47})--(\ref{51}) in the whole kinematical region (model A). We have also used the pole dependence for form factors $f_+(q^2)$, $A_2(q^2)$, $V(q^2)$ and $A_1(q^2)=\frac{2M_B M_\rho}{(M_B+M_\rho)^2}(1+w) \frac{A_1(0)}{ 1-q^2/m_P^2}$ (model B), which corresponds to replacing the function $\tilde\xi(w)$ (\ref{51}) by the pole form factor. The results are presented in Table 2 in comparison with the quark model \cite{16,17}, QCD sum rule \cite{18,19} and lattice (for $B\to\pi e\nu$) \cite{l1} predictions. Lattice accuracy, at present, is not enough to estimate $B\to\rho e\nu$ rates \cite{l3}. We see that our results for the above mentioned models A and B of form factor $q^2$-dependence coincide within errors. The ratio of the rates $\Gamma(B\to \rho e\nu)/\Gamma(B\to \pi e\nu)$ is considerably reduced in our model compared to the BSW \cite{16} and ISGW \cite{17} models with the simple pole or exponential $q^2$-behaviour of all form factors. Meanwhile our prediction for this ratio is in agreement with QCD sum rule results \cite{18,19}. The absolute values of the rates $\Gamma(B\to\pi e\nu)$ and $\Gamma(B\to\rho e\nu)$ in our model are close to those from QCD sum rules \cite{19}. The predictions for the rates with longitudinally and transversely polarized $\rho$ meson differ considerably in these approaches. This is mainly due to different $q^2$-behaviour of $A_1$ (see (\ref{48}), (\ref{52}) or pole dominance model \cite{16}). Thus the measurement of the ratios $\Gamma(B \to\rho e\nu)/\Gamma(B\to\pi e\nu)$ and $\Gamma_L/\Gamma_T$ should provide the test of $q^2$-dependence of $A_1$ and may discriminate between these approaches. The differential decay spectra $\frac{1}{\Gamma}\frac{d\Gamma}{d x}$ for $B\to \pi(\rho)$ semileptonic transitions, where $x=\frac{E_l}{M_B}$ and $E_l$ is the lepton energy, are presented in ref.~\cite{10a} (see Fig.~3). We can use our results for $V$ and $A_1$ to test the HQET relation \cite{20} between the form factors of the semileptonic and rare radiative decays of $B$ mesons. Isgur and Wise \cite{20} have shown that in the limit of infinitely heavy $b$-quark mass an exact relation connects the form factors $V$ and $A_1$ with the rare radiative decay $B\to\rho\gamma$ form factor $F_1$ defined by: \begin{eqnarray} \label{53} \langle \rho(p_\rho,e)\vert \bar ui\sigma _{\mu \nu}q^\nu P_Rb\vert B(p_B)\rangle&=&i\epsilon_{\mu \nu \tau \sigma }e^{*\nu }p_B^\tau p_\rho^\sigma F_1(q^2)\nonumber\\ & &+\big[e_\mu ^*(M_B^2-M_\rho^2)-(e^* q)(p_B+p_\rho)_\mu\big]G_2(q^2).\end{eqnarray} This relation is valid for $q^2$ values sufficiently close to $q^2_{max}=(M_B-M_\rho)^2$ and reads: \begin{equation} \label{54} F_1(q^2)=\frac{q^2+M_B^2-M_\rho^2}{ 2M_B} \frac{V(q^2)}{ M_B+M_\rho}+\frac{M_B+M_\rho}{2M_B}A_1(q^2).\end{equation} It has been argued in \cite{21,22,19}, that in these processes the soft contributions dominate over the hard perturbative ones, and thus the Isgur-Wise relations (\ref{54}) could be extended to the whole range of $q^2$. In \cite{10} we developed $1/m_b$ expansion for the rare radiative decay form factor $F_1(0)$ using the same ideas as in the present discussion of semileptonic decays. It was shown that Isgur-Wise relation (\ref{54}) is satisfied in our model at leading order of $1/m_b$ expansion. The found value of the form factor of rare radiative decay $B\to\rho\gamma$ up to the second order in $1/m_b$ expansion is \cite{10} \begin{equation} \label{55} F_1^{B\to\rho}(0)=0.26\pm 0.03.\end{equation} Using (\ref{54}) and the values of form factors (\ref{46}) we find \begin{equation} \label{56} F_1^{B\to\rho}(0)=0.27\pm 0.03,\end{equation} which is in accord with (\ref{55}). Thus we conclude that $1/m_b$ and $1/m_b^2$ corrections do not break the Isgur-Wise relation (\ref{54}) in our model. \section{CONCLUSIONS} We have presented in detail the method of the calculation of electroweak decay matrix elements for transitions between different meson states, based on the quasipotential approach in quantum field theory. It has been shown that the heavy-to-heavy decay matrix element can be expanded in inverse powers of the initial and final heavy quark masses at the point of zero recoil of the final meson. On the other hand, the heavy-to-light decay matrix element can be expanded in inverse powers of the initial heavy quark mass and large recoil momentum of final light meson at the point of maximum recoil of final meson. As a result the expansion of the heavy-to-light decay matrix element in inverse powers of initial heavy quark mass arises. This method permits the calculation of various radiative and weak decays of heavy mesons with the complete account of relativistic effects. This method has been applied to the investigation of the semileptonic decays of $B$ mesons into light mesons. The recoil momentum of final $\pi(\rho)$ meson is large compared to the $\pi(\rho)$ mass almost in the whole kinematical range. This requires the completely relativistic treatment of these decays. On the other hand, the presence of large recoil momentum, which for $q^2=0$ is of order $m_b/2$, allows for the $1/m_b$ expansion of weak decay matrix element at this point. The contributions to this expansion come both from the heavy $b$-quark mass and large recoil momentum of the light final meson. We have performed the $1/m_b$ expansion of the semileptonic decay form factors at $q^2=0$ up to the second order. The $q^2$-dependence of the form factors near $q^2=0$ has been determined. It has been found that the axial form factor $A_1$ has a $q^2$-behaviour different from other form factors (see (\ref{47})--(\ref{50})). This is in agreement with recent QCD sum rule results \cite{18,19,19a}. The ratios $\Gamma(B\to\rho e\nu)/\Gamma(B\to\pi e\nu)$ and $\Gamma_L/\Gamma_T$ are very sensitive to the $q^2$-dependence of $A_1$, and thus their experimental measurement may discriminate between different approaches. We have considered the relation between semileptonic decay form factors and the rare radiative decay form factor \cite{20}, obtained in the limit of the infinitely heavy $b$-quark. It has been found that in our model $1/m_b$ corrections do not violate this relation. \bigskip \noindent {\bf ACKNOWLEDGEMENTS} \nopagebreak \medskip \noindent We express our gratitude to B.~A.~Arbuzov, M.~A.~Ivanov, J.~G.~K\"orner, V.~A.~Matveev, M.~Neubert, V.~I.~Savrin, B.~Stech, A. Vainshtein for the interest in our work and helpful discussions of the results. This research was supported in part by the Russian Foundation for Fundamental Research under Grant No.94-02-03300-a and by the Interregional Centre for Advanced Studies. \begin{appendix} \section{ APPENDIX: HEAVY-TO-LIGHT SEMILEPTONIC DECAY FORM FACTORS AT $q^2=0$ POINT} \noindent a) $B\to\pi e\nu$ decay form factor \nopagebreak \begin{eqnarray}\label{a1} f_+^{(1)}(0)&=&\sqrt{\frac{E_\pi}{ M_B}} \int\frac{d^3p}{ (2\pi)^3} \bar\Psi_\pi\left({\bf p}+\frac{2\varepsilon_q}{ E_\pi+M_\pi}{\bf\Delta}_{max}\right) \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & &\times \sqrt{\frac{\varepsilon_b(p)+m_b}{2m_b}} \biggl\{1+\frac{M_B-E_\pi}{\varepsilon_q(p+\Delta_{max})+m_q} +\frac{({\bf p\Delta}_{max})}{{\bf\Delta}_{max}^2} \biggl(\frac{\varepsilon_q(p +\Delta_{max})-m_q}{2m_b}\nonumber\\ & & + (M_B-E_\pi)\left(\frac{1}{\varepsilon_q(p +\Delta_{max})+m_q}+ \frac{1}{2m_b}\right)\biggr) +(p_x^2+p_y^2)\Biggl(\frac{E_\pi-M_\pi}{2m_b(\varepsilon_q(p+ \Delta_{max}) +m_q)}\nonumber \\ & &\times \left( \frac{1}{\varepsilon_q(p)+m_q} -\frac{1}{\varepsilon_q(p+\Delta_{max})+m_q}\right) +\frac{M_B-E_\pi}{E_\pi+M_\pi}\left(\frac{1}{\varepsilon_q(p +\Delta_{max}) +m_q} -\frac{1}{2m_b}\right)\nonumber\\ & &\times \left( \frac{1}{\varepsilon_q(p)+m_q} -\frac{1}{\varepsilon_q(p+\Delta_{max}) +m_q}\right)\biggr)\Biggr\} \Psi_B({\bf p}),\\ \label{a2} f_+^{S(2)}(0)&=&\sqrt{\frac{E_\pi}{ M_B}} \int\frac{d^3p}{ (2\pi)^3} \bar\Psi_\pi\left({\bf p}+\frac{2\varepsilon_q}{ E_\pi+M_\pi}{\bf\Delta}_{max}\right) \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & &\times \Biggl\{\left(\frac{\varepsilon_q(\Delta_{max})-m_q}{ \varepsilon_q(\Delta_{max})+m_q}- \frac{M_B-E_\pi}{\varepsilon_q(\Delta_{max})+m_q}\right) \frac{1}{\varepsilon_q( \Delta_{max})}\left(M_\pi-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{E_\pi+M_\pi}\Delta_{max}\right)\right)\nonumber\\ & & +\frac{({\bf p\Delta}_{max})}{2{\bf \Delta}_{max}^2}\biggl[\biggl( \frac{\varepsilon_q(\Delta_{max})-m_q}{\varepsilon_q(\Delta_{max}) (\varepsilon_q(\Delta_{max})+m_q)} \nonumber \\ & & - (M_B-E_\pi)\left(\frac{1}{ \varepsilon_q(\Delta_{max})(\varepsilon_q(\Delta_{max})+m_q)}+ \frac{1}{m_b(\varepsilon_b(\Delta_{max})+m_b)}\right)\biggr) \nonumber \\ & & \times\left(M_B+M_\pi-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\pi+M_\pi}\Delta_{max}\right)\right)\nonumber \\ & & - \frac{\varepsilon_q(\Delta_{max})-m_q}{2m_b(\varepsilon_b( \Delta_{max})+m_b)}(M_B-\varepsilon_b(p)-\varepsilon_q(p)) +\frac{M_B-E_\pi}{2m_b\varepsilon_q(\Delta_{max})}\nonumber \\ & & \times\frac{\varepsilon_q(\Delta_{max})- m_q}{\varepsilon_q(\Delta_{max}) +m_q} \left(M_\pi-2\varepsilon_q\left(p+\frac{2\varepsilon_q}{E_\pi +M_\pi}\Delta_{max}\right)\right)\biggr]\nonumber\\ & & +\frac{p_x^2+p_y^2}{2(\varepsilon_q(p) +m_q)}\biggl[\biggl( \frac{E_\pi-M_\pi}{\varepsilon_q(\Delta_{max})(\varepsilon_q(p+ \Delta_{max})+m_q)^2} \nonumber\\ & & -\frac{M_B-E_\pi}{E_\pi+M_\pi}\biggl( \frac{1}{\varepsilon_q(\Delta_{max})(\varepsilon_q(\Delta_{max})+m_q)} - \frac{1}{m_b(\varepsilon_b(\Delta_{max})+m_b)}\biggr)\biggr) \nonumber\\ & &\times\left(M_B+M_\pi-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\pi+M_\pi}\Delta_{max}\right)\right)\nonumber\\ & & +\frac{(E_\pi-M_\pi)(M_B-\varepsilon_b(p)-\varepsilon_q(p))}{ m_b(\varepsilon_b(\Delta_{max})+m_b)(\varepsilon_q(\Delta_{max}) +m_q)}- \frac{M_B-E_\pi}{(E_\pi+M_\pi)m_b\varepsilon_q(\Delta_{max})} \nonumber\\ & &\times\frac{\varepsilon_q(\Delta_{max})-m_q} {\varepsilon_q(\Delta_{max})+m_q}\left(M_\pi- 2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\pi+M_\pi}\Delta_{max}\right)\right)\biggr]\Biggr\} \Psi_B({\bf p}),\\ \label{a3} f_+^{V(2)}(0)&=&\sqrt{\frac{E_\pi}{ M_B}} \int\frac{d^3p}{ (2\pi)^3} \bar\Psi_\pi\left({\bf p}+\frac{2\varepsilon_q}{ E_\pi+M_\pi}{\bf\Delta}_{max}\right) \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & & \times \Biggr\{-\frac{p_x^2+p_y^2}{(\varepsilon_q(p) +m_q)}\left(\frac{1}{\varepsilon_q(p)+m_q} -\frac{1}{\varepsilon_q(\Delta_{max})+m_q}\right)\nonumber\\ & & \times \biggl(\frac{E_\pi -M_\pi}{\varepsilon_q(p+ \Delta_{max})+m_q} \biggl(\frac{1}{\varepsilon_b(\Delta_{max}) +m_b} +\frac{1}{\varepsilon_q(\Delta_{max})+m_q}\biggr)\nonumber \\ & &-\frac{M_B-E_\pi}{E_\pi+M_\pi}\biggl( \frac{1}{\varepsilon_q(\Delta_{max})+m_q} - \frac{1}{\varepsilon_b(\Delta_{max})+m_b}\biggr)\biggr) \nonumber\\ & &\times\left(M_B+M_\pi-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\pi+M_\pi}\Delta_{max}\right)\right)\nonumber\\ & & +\frac{({\bf p\Delta}_{max})}{{\bf \Delta}_{max}^2}\frac{1}{ \varepsilon_q(p)+m_q} \biggl[ (\varepsilon_q(\Delta_{max}) -m_q) \nonumber\\ & & \times \left(\frac{1}{\varepsilon_b(\Delta_{max})+m_b}- \frac{m_q} {\varepsilon_q(\Delta_{max}) (\varepsilon_q(\Delta_{max})+m_q)} \right)\nonumber \\ & & + (M_B-E_\pi)\left(\frac{1}{ \varepsilon_q(\Delta_{max})+m_q} +\frac{1}{\varepsilon_b(\Delta_{max})+m_b}\right)\biggr] \nonumber \\ & & \times\left(M_B+M_\pi-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\pi+M_\pi}\Delta_{max}\right)\right)\nonumber \\ & & +\frac{{\bf p}^2}{\varepsilon_q(p)+m_q}\biggl[ \frac{M_B-M_\pi-\varepsilon_b(p)- \varepsilon_q(p)+2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\pi+M_\pi}\Delta_{max}\right)}{2\varepsilon_q(\Delta_{max}) (\varepsilon_q(\Delta_{max})+m_q)} \nonumber \\ & & -\frac{M_B-E_\pi}{\varepsilon_q(\Delta_{max})+m_q} \frac{M_B+M_\pi-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\pi+M_\pi}\Delta_{max}\right)}{2\varepsilon_q(\Delta_{max}) (\varepsilon_q(\Delta_{max})+m_q)} \nonumber \\ & & +\left(\frac{1}{\varepsilon_q(\Delta_{max})+m_q} -\frac{1}{2m_b} \left(1-\frac{M_B-E_\pi}{\varepsilon_q(\Delta_{max})+m_q}\right) \right)\frac{M_B-\varepsilon_b(p)-\varepsilon_q(p)}{\varepsilon_b( \Delta_{max})+m_b} \biggr]\Biggr\}\Psi_B({\bf p}), \end{eqnarray} \noindent b) $B\to \rho e\nu$ decay form factors \nopagebreak \begin{eqnarray}\label{a4} A_1^{(1)}(0)&=&\frac{2\sqrt{M_BM_\rho}}{ M_B+M_\rho} \sqrt{\frac{E_\rho}{ M_\rho}} \int\frac{d^3p}{ (2\pi)^3} \bar\Psi_\rho\left({\bf p}+\frac{2\varepsilon_q}{ E_\rho+M_\rho}{\bf\Delta}_{max}\right) \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & &\times \sqrt{\frac{\varepsilon_b(p)+m_b}{2m_b}} \Biggl\{1+\frac{1}{2m_b(\varepsilon_q(p+\Delta_{max})+m_q)} \biggl((E_\rho-M_\rho)\frac{p_x^2+p_y^2}{\varepsilon_q(p)+m_q} \nonumber\\ & & -\frac{1}{3}{\bf p}^2 -({\bf p\Delta}_{max})\biggr)\Biggr\} \Psi_B({\bf p}),\\ \label{a5} A_1^{S(2)}(0)&=&\frac{2\sqrt{M_BM_\rho}}{ M_B+M_\rho} \sqrt{\frac{E_\rho}{ M_\rho}} \int\frac{d^3p}{ (2\pi)^3} \bar\Psi_\rho\left({\bf p}+\frac{2\varepsilon_q}{ E_\rho+M_\rho}{\bf\Delta}_{max}\right) \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & &\times \Biggl\{\frac{\varepsilon_q(\Delta_{max})-m_q}{ \varepsilon_q(\Delta_{max})+m_q}\frac{1}{\varepsilon_q(\Delta_{max})} \left(M_\rho -2\varepsilon_q\left(p+\frac{2\varepsilon_q}{E_\rho+M_\rho}\Delta_{max} \right)\right) \nonumber\\ & & -\frac{(p_x^2+p_y^2)(E_\rho-M_\rho)}{2m_b(\varepsilon_q(p)+m_q) (\varepsilon_q(\Delta_{max})+m_q)} \biggl[\frac{1}{\varepsilon_b(\Delta_{max})+m_b}\nonumber\\ & & \times\left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\nonumber\\ & & +\frac{1}{\varepsilon_q(\Delta_{max})}(M_B-\varepsilon_b(p) -\varepsilon_q(p))\biggr] + \frac{({\bf p\Delta}_{max})} {{\bf\Delta}_{max}^2} \frac{\varepsilon_q(\Delta_{max})-m_q}{2} \biggl[\biggl( \frac{1}{m_b(\varepsilon_b(\Delta_{max})+m_b)} \nonumber\\ & & +\frac{1}{\varepsilon_q(\Delta_{max})(\varepsilon_q(\Delta_{max}) +m_q)}\biggr) \left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\nonumber\\ & &+\frac{1}{m_b\varepsilon_q(\Delta_{max})}(M_B-\varepsilon_b(p)- \varepsilon_q(p))\biggr]\Biggr\} \Psi_B({\bf p}),\\ \label{a6} A_1^{V(2)}(0)&=&\frac{2\sqrt{M_BM_\rho}}{ M_B+M_\rho} \sqrt{\frac{E_\rho}{ M_\rho}} \int\frac{d^3p}{ (2\pi)^3} \bar\Psi_\rho\left({\bf p}+\frac{2\varepsilon_q}{ E_\rho+M_\rho}{\bf\Delta}_{max}\right) \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & &\times \Biggl\{\frac{(p_x^2+p_y^2)(E_\rho-M_\rho)} {(\varepsilon_q(p)+m_q)(\varepsilon_q(\Delta_{max})+m_q)} \biggl(\frac{1}{(\varepsilon_q(p)+m_q)(\varepsilon_b(\Delta_{max}) +m_b)} \nonumber\\ & &+\frac{1}{(\varepsilon_q(\Delta_{max})+m_q)^2}\biggr) \left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\nonumber\\ & & +\frac{{\bf p}^2}{\varepsilon_q(p)+m_q}\biggl[ \frac{1}{2\varepsilon_q (\Delta_{max})(\varepsilon_q(\Delta_{max})+m_q)} \biggl( \frac{\varepsilon_q(\Delta_{max})- m_q}{\varepsilon_q(\Delta_{max})+m_q} \nonumber\\ & &\times\biggl(M_\rho -2\varepsilon_q\left(p+\frac{2\varepsilon_q}{E_\rho+ M_\rho}\Delta_{max} \right)\biggr) -(M_B-\varepsilon_b(p)-\varepsilon_q(p))\biggr)\nonumber\\ & &-\frac{1} {\varepsilon_b(\Delta_{max})+m_b}\left(\frac{1}{3(\varepsilon_q( \Delta_{max})+m_q)}+\frac{1}{m_b}\right)\nonumber\\ & &\times\left(M_\rho- 2\varepsilon\left(p+\frac{2\varepsilon_q}{E_\rho+ M_\rho}\Delta_{max} \right)\right)\biggr]- \frac{({\bf p\Delta}_{max})} {{\bf\Delta}_{max}^2} (\varepsilon_q(\Delta_{max})-m_q) \nonumber \\ & &\times\biggl[\frac{1}{2\varepsilon_q(\Delta_{max})(\varepsilon_q (\Delta_{max})+m_q)}\biggl(2(M_B-\varepsilon_b(p)-\varepsilon_q(p)) \nonumber\\ & &+\frac{\varepsilon_q(\Delta_{max})-m_q}{\varepsilon_q(\Delta_{max}) +m_q}\left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\biggr)\nonumber\\ & & +\frac{1}{\varepsilon_q(p)+m_q}\left( \frac{1}{\varepsilon_b(\Delta_{max})+m_b}+ \frac{m_q}{\varepsilon_q(\Delta_{max})(\varepsilon_q(\Delta_{max}) +m_q)}\right) \nonumber\\ & & \times \left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\biggr]\Biggr\} \Psi_B({\bf p}), \end{eqnarray} \begin{eqnarray} \label{a7} A_2^{(1)}(0)&=&\frac{M_B+M_\rho}{2\sqrt{M_BM_\rho}} \frac{2\sqrt{E_\rho M_\rho}}{E_\rho+ M_\rho} \int\frac{d^3p}{ (2\pi)^3} \bar\Psi_\rho\left({\bf p}+\frac{2\varepsilon_q}{ E_\rho+M_\rho}{\bf\Delta}_{max}\right) \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & & \times \sqrt{\frac{\varepsilon_b(p)+m_b}{2m_b}} \Biggl\{1+\frac{M_\rho}{M_B}\left(1-\frac{E_\rho+M_\rho} {\varepsilon_q(p+\Delta_{max})+m_q}\right) -\frac{p_z^2}{2m_b (\varepsilon_q(p+\Delta_{max})+m_q)}\nonumber\\ & & +\frac{p_x^2+p_y^2}{(\varepsilon_q(p+\Delta_{max})+m_q) (\varepsilon_q(p)+m_q)}\left(\frac{E_\rho+M_\rho}{2m_b} +\frac{M_\rho}{M_B}\right) \nonumber\\ & & -\frac{({\bf p\Delta}_{max})}{{\bf\Delta}_{max}^2}\left( \frac{\varepsilon_q(p+\Delta_{max})-m_q}{2m_b} +\frac{M_\rho}{M_B} \left(1+\frac{E_\rho}{m_b}\right)\right)\Biggr\} \Psi_B({\bf p}),\\ \label{a8} A_2^{S(2)}(0)&=&\frac{M_B+M_\rho}{2\sqrt{M_BM_\rho}} \frac{2\sqrt{E_\rho M_\rho}}{E_\rho+ M_\rho} \int\frac{d^3p}{ (2\pi)^3} \bar\Psi_\rho\left({\bf p}+\frac{2\varepsilon_q}{ E_\rho+M_\rho}{\bf\Delta}_{max}\right) \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & &\times \Biggl\{\left(\frac{\varepsilon_q(\Delta_{max})-m_q}{ \varepsilon_q(\Delta_{max})+m_q}+2\frac{M_\rho}{M_B}\right) \frac{1}{\varepsilon_q(\Delta_{max})} \left(M_\rho -2\varepsilon_q\left(p+\frac{2\varepsilon_q}{E_\rho+M_\rho}\Delta_{max} \right)\right) \nonumber\\ & & -\frac{p_x^2+p_y^2}{2m_b(\varepsilon_q(p)+m_q)} \biggl[\frac{1}{\varepsilon_b(\Delta_{max})+m_b}\nonumber\\ & & \times\left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\nonumber\\ & & +\frac{1}{\varepsilon_q(\Delta_{max})}(M_B-\varepsilon_b(p) -\varepsilon_q(p))\biggr] + \frac{({\bf p\Delta}_{max})} {{\bf\Delta}_{max}^2} \frac{1}{2}\biggl[\biggl(\frac{ \varepsilon_q(\Delta_{max})}{m_b(\varepsilon_b(\Delta_{max})+m_b)} \nonumber\\ & &+\frac{1}{\varepsilon_q(\Delta_{max})+m_q}\biggr) \left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\nonumber\\ & &+\frac{1}{m_b}(M_B-\varepsilon_b(p)- \varepsilon_q(p))\biggr]\Biggr\} \Psi_B({\bf p}),\\ \label{a9} A_2^{V(2)}(0)&=&\frac{M_B+M_\rho}{2\sqrt{M_BM_\rho}} \frac{2\sqrt{E_\rho M_\rho}}{E_\rho+ M_\rho} \int\frac{d^3p}{ (2\pi)^3} \bar\Psi_\rho\left({\bf p}+\frac{2\varepsilon_q}{ E_\rho+M_\rho}{\bf\Delta}_{max}\right) \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & &\times \Biggl\{\frac{p_x^2+p_y^2} {(\varepsilon_q(p)+m_q)(\varepsilon_q(\Delta_{max})+m_q)} \biggl(\frac{E_\rho+M_\rho}{(\varepsilon_q(p)+m_q)(\varepsilon_b( \Delta_{max}) +m_b)} \nonumber\\ & &+\frac{1}{\varepsilon_q(\Delta_{max})+m_q}\biggr) \left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\nonumber\\ & & -\frac{{\bf p}^2}{\varepsilon_q(p)+m_q}\biggl[ \frac{1}{2\varepsilon_q (\Delta_{max})(\varepsilon_q(\Delta_{max})+m_q)} \nonumber\\ & &\times\biggl(M_B+M_\rho-\varepsilon_b(p)-\varepsilon_q(p) -2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{E_\rho+ M_\rho}\Delta_{max} \right)\biggr) \nonumber\\ & &+\frac{1}{\varepsilon_b(\Delta_{max})+m_b} \left(\frac{1}{3(\varepsilon_q(\Delta_{max})+m_q)} +\frac{1}{m_b}\right)\left(M_\rho-2\varepsilon\left(p +\frac{2\varepsilon_q}{E_\rho+ M_\rho}\Delta_{max} \right)\right)\biggr]\nonumber\\ & &- \frac{({\bf p\Delta}_{max})}{{\bf\Delta}_{max}^2} \biggl[\frac{1}{\varepsilon_q(p)+m_q} \left( \frac{\varepsilon_q(\Delta_{max})+m_q}{\varepsilon_b(\Delta_{max}) +m_b}+\frac{m_q}{2\varepsilon_q(\Delta_{max})} \right)\nonumber\\ & &\times \left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\biggr)\nonumber\\ & & +\frac{1}{\varepsilon_q(\Delta_{max})}\biggl(3(M_B- \varepsilon_b(p) -\varepsilon_q(p))\nonumber\\ & &-\left(M_\rho-2\varepsilon_q\left(p+\frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\biggr)\biggr]\Biggr\} \Psi_B({\bf p}),\end{eqnarray} \begin{eqnarray} \label{a10} V^{(1)}(0)&=&\frac{M_B+M_\rho}{2\sqrt{M_BM_\rho}} \int\frac{d^3p}{(2\pi)^3} \bar\Psi_\rho\left({\bf p} +\frac{2\varepsilon_q}{ E_\rho+M_\rho}{\bf\Delta}_{max}\right) \frac{2\sqrt{E_\rho M_\rho}}{\varepsilon_q(p+\Delta_{max})+m_q} \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & & \times \sqrt{\frac{\varepsilon_b(p)+m_b}{2m_b}} \Biggl\{1+\frac{p_x^2+p_y^2}{E_\rho+M_\rho}\left( \frac{\varepsilon_q(p+\Delta_{max})+m_q}{2m_b(\varepsilon_q(p)+m_q)} -\frac{1}{\varepsilon_q(p+\Delta_{max})+m_q}\right)\nonumber\\ & & +\frac{({\bf p\Delta}_{max})}{{\bf\Delta}_{max}^2}\left( 1- \frac{\varepsilon_q(p+\Delta_{max})+m_q}{2m_b} \right) \Biggr\} \Psi_B({\bf p}),\\ \label{a11} V^{S(2)}(0)&=&\frac{M_B+M_\rho}{2\sqrt{M_BM_\rho}} \int\frac{d^3p}{(2\pi)^3} \bar\Psi_\rho\left({\bf p} +\frac{2\varepsilon_q}{ E_\rho+M_\rho}{\bf\Delta}_{max}\right) \frac{2\sqrt{E_\rho M_\rho}}{\varepsilon_q(p+\Delta_{max})+m_q} \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & &\times \Biggl\{-\frac{1}{\varepsilon_q(\Delta_{max})} \left(M_\rho -2\varepsilon_q\left(p+\frac{2\varepsilon_q} {E_\rho+M_\rho}\Delta_{max} \right)\right) +\frac{p_x^2+p_y^2}{2m_b(E_\rho+M_\rho)(\varepsilon_q(p)+m_q)} \nonumber\\ & & \times\biggl[\frac{\varepsilon_q(p+\Delta_{max})-m_q} {\varepsilon_q(\Delta_{max})} \left(M_\rho -2\varepsilon_q\left(p+\frac{2\varepsilon_q} {E_\rho+M_\rho}\Delta_{max} \right)\right) \nonumber\\ & & -\frac{\varepsilon_q(\Delta_{max})+m_q}{\varepsilon_b (\Delta_{max}) +m_b} \left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\biggr]\nonumber\\ & & - \frac{({\bf p\Delta}_{max})} {{\bf\Delta}_{max}^2} \biggl[\frac{1}{2}\left(\frac{1}{\varepsilon_q(\Delta_{max})} - \frac{\varepsilon_q(\Delta_{max})+m_q}{ m_b (\varepsilon_b(\Delta_{max}) +m_b)}\right)\nonumber\\ & &\times \left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\nonumber\\ & &-\frac{\varepsilon_q(\Delta_{max})-m_q}{2m_b \varepsilon_q(\Delta_{max})} \left(M_\rho -2\varepsilon_q\left(p+\frac{2\varepsilon_q} {E_\rho+M_\rho}\Delta_{max} \right)\right) \biggr]\Biggr\} \Psi_B({\bf p}),\\ \label{a12} V^{V(2)}(0)&=&\frac{M_B+M_\rho}{2\sqrt{M_BM_\rho}} \int\frac{d^3p}{(2\pi)^3} \bar\Psi_\rho\left({\bf p} +\frac{2\varepsilon_q}{E_\rho+M_\rho}{\bf\Delta}_{max}\right) \frac{2\sqrt{E_\rho M_\rho}}{\varepsilon_q(p+\Delta_{max})+m_q} \sqrt{\frac{\varepsilon_q(p+\Delta_{max})+m_q}{ 2\varepsilon_q(p+\Delta_{max})}} \nonumber \\ & &\times \Biggl\{\frac{2(p_x^2+p_y^2)}{(E_\rho+M_\rho) (\varepsilon_q(p)+m_q)}\biggl(\frac{\varepsilon_q(\Delta_{max})+m_q} {(\varepsilon_q(p)+m_q)(\varepsilon_b(\Delta_{max})+m_b)} \nonumber\\ & &-\frac{1}{\varepsilon_q(\Delta_{max})+m_q}\biggr) \left(M_B+M_\rho-\varepsilon_b(p)- \varepsilon_q(p)-2\varepsilon_q\left(p+ \frac{2\varepsilon_q}{ E_\rho+M_\rho}\Delta_{max}\right)\right)\nonumber\\ & & -\frac{{\bf p}^2}{2m_b(\varepsilon_q(p)+m_q)}\biggl[ \frac{1}{\varepsilon_b(\Delta_{max})+m_b}(M_B-\varepsilon_b(p) -\varepsilon_q(p))\nonumber\\ & & +\frac{1}{\varepsilon_q(\Delta_{max})+m_q} \left(M_B+M_\rho-\varepsilon_b(p)-\varepsilon_q(p)- 2\varepsilon\left(p+\frac{2\varepsilon_q}{E_\rho+ M_\rho}\Delta_{max} \right)\right)\biggr] \nonumber\\ & & + \frac{({\bf p\Delta}_{max})} {{\bf\Delta}_{max}^2} \frac{1}{\varepsilon_q(p)+m_q}\left(1- \frac{\varepsilon_q(\Delta_{max})+m_q}{\varepsilon_b (\Delta_{max})+m_b}\right)\nonumber\\ & &\times\left(M_B+M_\rho-\varepsilon_b(p)-\varepsilon_q(p)- 2\varepsilon\left(p+\frac{2\varepsilon_q}{E_\rho+ M_\rho}\Delta_{max} \right)\right)\Biggr\} \Psi_B({\bf p}), \end{eqnarray} where the superscripts ``(1)" and ``(2)" correspond to Figs.~1 and 2, $S$ and $V$ --- to the scalar and vector potentials of $q\bar q$-interaction; \begin{equation}\label{a13}\vert {\bf\Delta}_{max}\vert=\frac{M_B^2- M_{\pi(\rho)}^2}{ 2M_B};\qquad E_{\pi(\rho)}=\sqrt{M_{\pi(\rho)}^2+ {\bf\Delta}_{max}^2}= \frac{M_B^2+ M_{\pi(\rho)}^2}{ 2M_B}; \end{equation} and $z$-axis is chosen in the direction of ${\bf\Delta}_{max}$. \end{appendix} \frenchspacing
\section{Introduction} The contouring methods described by \CITEX{Lewis-MiraldaEscude-Rich-Wambs93} and \CITEX{Witt93-contouring-APJ} are very efficient for obtaining the magnification of a point source moving along a straight track in the source plane. For finite sources, however, the amplification {\em must} be computed for numerous parallel tracks and then convolved with the source profile. Rayshooting, on the other hand, is an efficient algorithm for relatively large sources, but the computing time increases with the inverse of the source area for a given noise level. \section{The hybrid method} \label{sec:method} By using the method described in \CITEX{Lewis-MiraldaEscude-Rich-Wambs93}, all the images of a straight, infinite line in the source plane can be found. The images are the borders between those parts of the lens plane projected above the straight line, and those parts projected below the straight line. After finding the images of one line below the source and one line above the source, it is clear that those parts of the lens plane that are projected between the two infinite lines in the source plane are the areas between the images of the infinite lines. Furthermore, those segments corresponding to the upper and lower edges of a box surrounding the source may be identified. The end points of these segments are projected onto the corners of the ``source box''. Starting from the corner points, the contouring method can be ``turned around'' 90 degrees, and all the lines joining all the corner points of the ``source box'' are found. After this step, all the images of the source box are placed within known, closed polygons. Rayshooting is then performed within all the closed polygons, and the lightcurve is produced in the usual way. \section{Efficiency} \label{sec:efficiency} The efficiency of the rayshooting part of the method compared to crude, non-optimized rayshooting can be found by comparing the size of the areas where rayshooting has to be performed. A target area in the source plane with length $2l$, and height $2r_{\rm s}$ gives an effective lightcurve length $L_{\rm c} = 2l-2r_{\rm s}$, where $r_{\rm s}$ is the source radius. The theoretical efficiency $f$ can be shown to be given by \begin{eqnarray} f &\approx& \left\{ \begin{array}[c]{l} (1 + \bfrac{10\sqrt{\kappa_*}}{r_{\rm s}} + % {} \bfrac{100\kappa_*}{lr_{\rm s}}) \\[.5cm] (1+ \bfrac{20\sqrt{\kappa_*}}{r_{\rm s}} + \bfrac{100\kappa_*}{r_{\rm s}^2}) \end{array} \right. \begin{array}[l]{l} \mbox{For $l \gg r_{\rm s}$} \\[.5cm] \mbox{For $l = r_{\rm s}$, $L_{\rm c}=0$.} \end{array} \label{eq:ratio} \end{eqnarray} \section{Discussion} \label{sec:discussion} The above arguments give a theoretical efficiency factor on the order of $10^5$ for e.g. a snapshot of the source with $r_{\rm s}=0.01$, $l=r_{\rm s}$ and $\kappa_*=0.4$. However, the most time-consuming task for the hybrid method is going to be the contouring itself. For a snapshot like the example above, the contouring amounts to about $10^5$ shots \cite{Lewis-MiraldaEscude-Rich-Wambs93}. This must be compared with the total number of shots necessary to get a specific signal to noise ratio, generally about $10^3$ shots. The highest estimates of $f$ thus have to be lowered by roughly a factor of $100$, depending on the specific parameters $r_{\rm s}$, $\kappa_*$, $\gamma$, and $l$. Even so, the proposed hybrid method has the potential to be a very efficient workhorse for producing accurate model lightcurves for small but extended sources. \section*{Acknowledgments} The author would like to thank Rolf Stabell and Sjur Refsdal for comments during the preparation of this poster.
\section{Introduction} Nuclear reactions reveal various aspects of hadronic many-body problem as a function of the target and projectile combination, the incident energy and the angular momentum involved. In nucleon induced reactions, for example, the compound process is dominant in low energy region, while the pre-equilibrium and spallation processes become more likely as the incident energy increases. In heavy-ion collisions, we also have to introduce various models of different natures depending on each specific process. However, most of them are restricted to the specific energy regime or specific phenomenon and some of them have too many parameters to obtain a definite physical conclusion from the analysis. The main purpose of the series of our work is to develop a model which can describe the various aspects of nuclear reactions in an unified way. We try to seek a model with minimum number of parameters, a wide range of applicability, and a quantitative agreement with as many observables as possible. In addition to these requests, we demand the model to be so simple that one can run its computer code on work stations. In the heavy-ion physics, microscopic models, which describe reactions in terms of the dynamics of the interacting nucleons, are commonly used to extract the information of the nuclear matter under extreme conditions from the final observables. The most popular models of this type are the Boltzmann-Uehling-Uhlenbeck/Vlasov-Uehling-Uhlenbeck (BUU/VUU) \cite{vuu00}, the quantum molecular dynamics (QMD) \cite{qmd00}, and CASCADE type models \cite{cascade00,cascade01}. So far these microscopic models have shedded light on several exciting topics in heavy-ion physics, e.g. the multifragmentation, the flow of the nuclear matter, and the energetic particle production. However the parameters of the models such as the effective interaction, elastic and inelastic channels of $NN$ cross section, differ substantially from one model to the other even in the same type of model. Furthermore, these models have not been tested intensively in much simpler light-ion reactions except for an analysis of (p,xn) reaction carried out by Peilert et al \cite{peilert}. In their analysis, however, the lower part of neutron energy spectra cannot be treated, since statistical decay following the QMD process was not included. We thus start the series of our work from the analysis of the simplest type of the reactions, the $(N,xN')$ (nucleon in, nucleon out) reaction in this paper, aiming to establish an unified model for various nuclear reactions. In the subsequent works, we are planning to analyze $(N,x\pi)$, $(\pi,xN)$ reactions and heavy-ion reactions. We restrict our subject to the reactions of nucleon-nucleus, meson-nucleus and nucleus-nucleus collisions with energies well above Coulomb barrier up to several GeV/nucleon, where the classical treatment of the collisions is justified in a first-order approximation. We do not deal with the phenomena which are dominated by the quantum effects. In this energy regime, the whole reaction process could be divided into two parts, i.e. dynamical process and statistical process. These two processes are well separated in their time scales. In the dynamical process, the direct reactions, nonequilibrium reactions, and dynamical formation of highly-excited fragments take place during typical collision times of the order $10^{-22}$ sec. After that, the evaporation and fission decay, which we call the statistical process, occur in the longer time scale of the order $10^{-21}-10^{-15}$ sec. We thus employ a two step model, namely, we incorporate quantum molecular dynamics (QMD) for the dynamical process with statistical decay model (SDM). Similar hybrid models have been used in the analysis of the heavy-ion collisions \cite{maru92a,g-qmd,f-qmd}. In this paper, we define the basic ingredients of QMD plus SDM model and discuss how these two are combined without introducing any additional parameter. We then apply this model systematically to $(N,xN')$ reactions, and discuss which element in the model is crucial for describing these reactions and what is necessary to develop the model further. In Sec. II we describe the details of the QMD, the effective interaction, $NN$ elastic and inelastic cross sections, the relativistic corrections and the statistical decay model employed in our model. In Sec. III we compare the various double differential cross sections calculated by this model with the experimental data for proton induced reactions with incident energies from 100 MeV up to 3 GeV. We summarize and conclude this work in Sec. IV. \section{Description of the Basic Model} As we mentioned above, our basic model consists of two part, the quantum molecular dynamics (QMD) and the statistical decay model (SDM). The reason of employing the QMD model for the description of the dynamical processes is that the QMD can calculate the fragment formation in a natural and practical way. Though the QMD method is widely used in the study of nuclear fragmentation \cite{qmd00}, the details of the prescription differ from author to author. Aiming to establish a simple standard model, we will start from the standard type of QMD, taking into account of the relativistic kinematics and the relativistic correction for the effective interaction. Additionally, we treat the resonances of nucleon, $\Delta$ and $N^*(1440)$, and real pions with their isospin degrees of freedom in the equation of motion. For the statistical decay process, we use a simple prescription including only the light particle evaporation. \subsection{Quantum Molecular Dynamics} \subsubsection{Basic equations and effective interaction} The QMD method is a semi-classical simulation method in which each nucleon state is represented by a Gaussian wave function of width $L$, \begin{equation} \phi_i({\bf r}) = \frac{1}{(2\pi L)^{3/4}} \exp \left[ - \frac{({\bf r} - {\bf R}_i)^2}{4L} + \frac{i}{\hbar} {\bf r} \cdot {\bf P}_i \right], \end{equation} where ${\bf R}_i$ and ${\bf P}_i$ are the centers of position and momentum of $i$-th nucleon, respectively. The total wave function is assumed to be a direct product of these wave functions. Thus the one-body distribution function is obtained by the Wigner transform of the wave function, \begin{equation} f({\bf r},{\bf p}) = \sum_i { f_i({\bf r},{\bf p}) }, \label{f0} \end{equation} \begin{equation} f_i({\bf r},{\bf p}) = 8 \cdot \exp\left[-{({\bf r}-{\bf R}_i)^2\over 2L} -{2L({\bf p}-{\bf P}_i)^2\over \hbar^2}\right]. \label{fi} \end{equation} The time evolution of ${\bf R}_i$ and ${\bf P}_i$ is described by Newtonian equations and the stochastic two-body collision term. The Newtonian equations are derived on the basis of the time-dependent variational principle \cite{qmd00} as \begin{equation} \dot{{\bf R}}_i = \frac{\partial H}{\partial {\bf P}_i}, \;\;\;\; \dot{{\bf P}}_i = - \frac{\partial H}{\partial {\bf R}_i}, \label{newton00} \end{equation} where the Hamiltonian $H$ consists of the single-particle energy including the mass term and the energy of the two-body interaction. As for the effective interaction, we adopt the Skyrme type, the Coulomb, and the symmetry terms in this paper. By using the Gaussian function of nucleons [Eq.\ (\ref{fi})], we get \begin{eqnarray} H & = & \sum_i\;{ E_i } \nonumber \\ & & \; + \; {1\over 2}{A \over\rho_{0}}\sum_i<\rho_i> \; + \; {1\over 1+\tau}{B \over \rho_{0}^{\tau}} \sum_i<\rho_i>^{\tau} \nonumber \\ & & \; + \; {1\over 2}\sum_{i , j(\neq i)} c_{i} \, c_{j} \, {e^2\over|{\bf R}_i-{\bf R}_j|} \, \> {\rm erf}\left( |{\bf R}_i-{\bf R}_j|/\sqrt{4L} \right) \nonumber \\ & & \; + \; {C_{\rm s}\over 2\rho_0} \sum_{i , j(\neq i)} \, ( 1 - 2 | c_i - c_j | ) \; \rho_{ij}, \label{ham1} \end{eqnarray} with \begin{equation} E_i = \sqrt{ m_i^2 + {\bf P}_i^2} \;, \end{equation} where erf denotes the error function. In the above equation, $c_i$ is 1 for protons and 0 for neutrons, while $<\rho_i>$ is an overlap of density with other nucleons defined as \begin{eqnarray} <\rho_i> & \equiv & \sum_{j\neq i} \; \rho_{ij} \; \equiv \sum_{j\neq i} { \int { d{\bf r} \; \rho_i({\bf r}) \; \rho_j({\bf r}) }} \nonumber \\ & = & \sum_{j\neq i}{ (4\pi L)^{-3/2} \exp \left[ - ( {\bf R}_i - {\bf R}_j ) ^2 / 4L \right] }, \label{rhoij} \end{eqnarray} with \begin{eqnarray} \rho_i({\bf r}) & \equiv & \int \frac{d{\bf p}}{(2\pi \hbar)^3} \; f_i({\bf r},{\bf p}) \nonumber \\ & = & (2\pi L)^{-3/2} \exp \left[ - ( {\bf r} - {\bf R}_i ) ^2 / 2L \right]. \end{eqnarray} In this paper we use the parameters $A=-219.4$ MeV, $B=165.3$ MeV, and $\tau=4/3$ which yield a compressibility of $K=237.3$ MeV, saturation at $\rho = \rho_0 = 0.168 {\rm fm}^{-3}$ and a binding energy of 16 MeV per nucleon for infinite nuclear matter. The symmetry energy parameter $C_{\rm s}$ is chosen to be 25 MeV. The width of Gaussian $L$ is a parameter of the QMD and fixed as $L=2.0 \; {\rm fm}^2$ in this paper. \subsubsection{Two-body collision term} In addition to the Newtonian equation Eq.\ (\ref{newton00}), the time evolution of the system is affected by the two-body collision term. In the QMD method, the stochastic two-body collision process is introduced in a phenomenological way on the analogy of the test-particle calculation of the BUU collision term \cite{vuu00}. It includes the Pauli blocking factor $(1-f({\bf r},{\bf p},t))$, which is lacking in the cascade collision process \cite{cascade00,cascade01}. We follow basically the prescription of the two-body collision term used in the BUU calculation done by Wolf et al. \cite{wolf,wolf2}, and modify it to extend the energy range up to 3 GeV. We thus describe here the outline of the procedure of Ref.\ \cite{wolf,wolf2} and explain the extensions introduced in this paper. Further details of the collision term and the dynamics of $\Delta$'s, $N^*$'s, and pions discussed below can be found in Ref.\ \cite{wolf,wolf2}. It is assumed that two particles collide if their impact parameter defined in a covariant way is smaller than a given value $b_{{\rm max}}=\sqrt{\sigma/\pi}$ obtained from cross section~$\sigma$. The collisions are considered as instantaneous interaction and a collision event is specified by the two points in space-time where the collision happens. Therefore it is hard to retain the covariance, since one has to choose a common reference frame for the QMD calculations. Hence the average proper time of the collision points defined by each particle is used to determine the time step in which the collision happens. This collision prescription was checked for heavy-ion collisions from 400 MeV/nucleon to 2.1 GeV/nucleon, and it was found that the disturbance of the covariance was very small \cite{wolf}. In order to treat the reactions with high bombarding energies, we include in our QMD simulation the nucleons ($N$), deltas ($\Delta$(1232)), $N^*$(1440)'s and pions with their isospin degree of freedom. The $\Delta$'s and $N^*$'s are propagated in the same interactions as the nucleons except for the symmetry term, while pions feel only the Coulomb interaction. The creation and absorption of these particles are treated in the collision term. In the following, we list all channels included in the collision term, where $B$ denotes a baryon and $N$, more specifically, a nucleon: \begin{eqnarray} \begin{array}{llllllll} 1.\;& B_i & +\;& B_j & \to \;& B_i & +\;& B_j \\ 2. & N & + & N & \to & N & + & \Delta \\ 3. & N & + & \Delta & \to & N & + & N \\ 4. & N & + & N & \to & N & + & N^* \\ 5. & N & + & N^* & \to & N & + & N \\ 6. & N & + & \pi & \to & \Delta & & \\ 7. & N & + & \pi & \to & N^* & & \\ 8. & \Delta & + & \pi & \to & N^* & . & \\ \end{array} \label{channel1} \end{eqnarray} The channel 8 has been added to the prescription of Wolf et al. \cite{wolf}, which is the inverse process of the additional decay channel of $N^*$(1440)'s (cf. channel 11 below). We use the following parametrization for all baryon-baryon elastic cross section (channel 1 in Eq.\ (\ref{channel1})), \begin{equation} \sigma = \frac{C_1}{1+100 \sqrt{s} \; '} + C_2 \; ({\rm mb}), \label{signn1} \end{equation} with \begin{equation} \sqrt{s} \; ' = \max( 0, \sqrt{s} - M_i - M_j - {\rm cutoff} ) \; ({\rm GeV}), \end{equation} where cutoff is 0.02 GeV for nucleon-nucleon channel, while it is zero for the others. This is the conventional Cugnon parametrization form \cite{vuu00,cascade01}. We use this form up to $\sqrt{s} \; ' = 0.4286 \; ({\rm GeV})$, which corresponds to 1 GeV Lab energy for nucleon-nucleon case. Above 1 GeV, we parametrize the experimental $p$-$p$ and $p$-$n$ elastic cross section \cite{ptable,data1} as, \begin{eqnarray} \sigma = C_3 \left[ 1 - \frac{2}{\pi} \tan^{-1} \left( 1.5 \sqrt{s} \; ' - 0.8 \right) \right] + C_4 \; ({\rm mb}). \label{signn2} \end{eqnarray} In order to connect Eqs.\ (\ref{signn1}) and (\ref{signn2}) smoothly, we slightly modified the parameters of Cugnon \cite{vuu00,cascade01}. The actual values of the parameters $C_i$ in the above equations are listed in Table \ref{table1}. The angular distribution of the elastic channels is taken from the same form as Cugnon parametrization \cite{vuu00,cascade01}. For the production of baryonic resonances (channel 2 and 4 in Eq.\ (\ref{channel1})), we adopt the total cross section based on the parametrization of VerWest and Arndt \cite{west}, in which the pion cross sections are parametrized assuming the pions are produced through baryonic resonances. Their parametrization was performed by fitting the experimental data up to 1.5 GeV incident energy. In order to extend the energy range up to 3 GeV, we have modified the parameters in the following way. In the model of VerWest and Arndt, the cross sections are parametrized according to the initial and final total isospin $i$ and $f$ of the two-nucleon system \cite{west} as \begin{equation} \sigma_{if}(s) = \frac{\pi(\hbar c)^2}{2p^2} \; \alpha \left( \frac{p_r}{p_0}\right)^{\beta} \frac{m_0^2 \Gamma^2 (q/q_0)^3} {(s^* - m^2_0)^2 + m_0^2 \Gamma^2}, \label{wests} \end{equation} where \begin{eqnarray} p_0^2 & = & {1\over 4} s_0 - m^2_N , \;\;\;\;\; s_0 = (m_N+m_0)^2 , \nonumber \\ p_r^2(s) & = & \frac{[s-(m_N-\langle M \rangle )^2][s-(m_N+\langle M \rangle )^2]} {4s}, \\ q^2(s^*) & = & \frac{[s^*-(m_N-m_\pi )^2][s^*-(m_N+m_\pi )^2]} {4s^*}, \nonumber \\ s^* & = & \langle M \rangle^2 , \;\;\;\; q_0 = q(m_0^2), \nonumber \end{eqnarray} and $\langle M \rangle$ is the mean mass of the resonance \cite{west} obtained from a Breit-Wigner distribution with $M_0 = 1220$ MeV, $\Gamma_0 = 120$ MeV for the $\Delta$ and $M_0 = 1430$ MeV and $\Gamma_0 = 200$ MeV for the $N^*$. From these cross sections, we determine the production cross section of $\Delta$'s \cite{wolf} as \begin{eqnarray} \begin{array}{lllllllll} p & + & p & \to & n & + & \Delta^{++} & : & \sigma_{10}+{1\over 2}\sigma_{11} ,\\[0.5ex] p & + & p & \to & p & + & \Delta^{+} & : & {3\over 2}\sigma_{11} ,\\[0.5ex] n & + & p & \to & p & + & \Delta^{0} & : & {1\over 2}\sigma_{11}+{1\over 4}\sigma_{10} ,\\[0.5ex] n & + & p & \to & n & + & \Delta^{+} & : & {1\over 2}\sigma_{11}+{1\over 4}\sigma_{10} ,\\[0.5ex] n & + & n & \to & p & + & \Delta^{-} & : & \sigma_{10}+{1\over 2}\sigma_{11} ,\\[0.5ex] n & + & n & \to & n & + & \Delta^{0} & : & {3\over 2}\sigma_{11} . \end{array} \label{delta} \end{eqnarray} We have effectively included the cross section of the $\pi d$ final state, parametrized as $\sigma^d_{10}$ in \cite{west}, in the cross section of $\sigma_{10}$. We assume in this paper that the cross section $\sigma_{01}$ in \cite{west} contributes only to the $N^*$ production independently of the isospin components. Thus we rename $\sigma_{01}$ as $\sigma_{N^*}$, and the $N^*$ production cross sections are given by \begin{eqnarray} \begin{array}{lllllllll} p & + & p & \to & p & + & N^{*+} & : & {3\over 2}\sigma_{N^*} ,\\[0.5ex] n & + & p & \to & p & + & N^{*0} & : & {3\over 4}\sigma_{N^*} ,\\[0.5ex] n & + & p & \to & n & + & N^{*+} & : & {3\over 4}\sigma_{N^*} ,\\[0.5ex] n & + & n & \to & n & + & N^{*0} & : & {3\over 2}\sigma_{ N^*} .\\[0.5ex] \end{array} \label{nstar} \end{eqnarray} The new parameters in Eq.\ (\ref{wests}) are given in Table \ref{table3}. In order to determine these parameters and the parameters of elastic cross section in high energy part defined in Eq.~(\ref{signn2}), we fitted the experimental $p$-$p$ and $p$-$n$ cross sections \cite{ptable,data1}. In Fig.\ \ref{pptot}, we show the $p$-$p$ (a) and $p$-$n$ (b) total (solid line), elastic (long dashed line), and inelastic (dot-dashed line) cross sections. The inelastic cross section is the sum of the $\Delta$ (short dashed line) and $N^*$ (dotted line) production cross section, calculated by Eqs.~(\ref{signn1},\ref{signn2},\ref{wests},\ref{delta},\ref{nstar}). In the same figure, we show the corresponding experimental total (open circles), elastic (open triangles), and inelastic (open boxes) cross sections \cite{data1} with error bars. For the $p$-$p$ case, the present parametrization of elastic and inelastic cross sections fits to the data for whole energy range up to 3 GeV except for the some deviation around 1 GeV, which is due to the elastic cross section and does not affect the result. On the other hand, for $p$-$n$ case where only the total cross section is available in the data, we fitted it at the energy higher than 0.7 GeV up to 3 GeV, and adopt the Cugnon's type elastic cross section in the low momentum region instead of the free elastic cross section. In Fig.\ \ref{pppi} we show the pion cross section of $pn \to nn\pi^+ + pp\pi^-$ (solid line) obtained by our new parametrization of Table \ref{table3}. In the same figure, the gray bold line denotes the result of the original parametrization of VerWest and Arndt \cite{west}, while the experimental data are taken from Ref.\ \cite{data1}. By this new parameter set, our pion production cross section below 1.5 GeV does not differ so much from the original results of VerWest and Arndt \cite{west}, which are essentially the same as the data. However, above 1.5 GeV, our result fitted the experimental data, while the result obtained by extrapolating the original parametrization of VerWest and Arndt to higher energy shows a big bump, which has no experimental support. In the higher energy region, the role of $N^*$ becomes important. One of the good quantities which shows the characteristics of the higher resonances is the elementary two pion production cross section. In the present prescription it is described only by $\sigma_{N^*}$ combined with the decay modes of the resonances which will be mentioned below. For example, it is shown in our prescription that \begin{eqnarray} \sigma(pn \to pp \pi^0 \pi^-) & = & \frac{1}{15} \sigma_{N^*}, \nonumber \\ \sigma(pn \to pn \pi^+ \pi^-) & = & \frac{5}{12} \sigma_{N^*}, \label{twopi} \\ \sigma(pp \to pp \pi^+ \pi^-) & = & \frac{1}{3} \sigma_{N^*}. \nonumber \end{eqnarray} We thus plot the $\sigma_{N^*}$ in Fig.\ \ref{signn3} as well as the experimental two pion production cross sections \cite{data1} scaled by above factors. This figure shows that our parametrization fitted the gross features of the experimental data for the energy range up to 3 GeV. Although this parametrization should be modified if two delta production channel or the direct two pion decay of $N^{*}$ or higher resonances are included, the present prescription of the elementary inelastic channels could roughly reproduce the experimental single and two pion production cross sections for the energy range up to 3 GeV. We do not take into account the direct s-state pion production mechanism but all pions are assumed to be produced through baryonic resonances. The masses of the resonances are randomly distributed according to the Breit-Wigner distribution with the momentum-dependent width \cite{moniz}, i.e., \begin{equation} f(M) = \frac{0.25 \Gamma^2}{(M-M_r)^2 + 0.25 \Gamma^2}, \label{masdis1} \end{equation} with \begin{equation} \Gamma = \left( \frac{q}{q_r} \right)^3 \frac{M_r}{M} \left( \frac{v(q)}{v(q_r)} \right)^2 \Gamma_r, \label{masdis2} \end{equation} where $q$ denotes the c.m. momentum in the $\pi N$ channel, and index $r$ refers to the values at the mass $M_r$, and \begin{equation} v(q) = \frac{\beta_r^2}{\beta_r^2 + q^2}. \label{masdis3} \end{equation} We have applied this momentum dependent width not only to the $\Delta$-resonance but also to $N^*$(1440). The values of the parameters used in this paper are listed in Table~\ref{table4}. Another important ingredient of the resonance production (channel 2 and 4 in Eq.\ (\ref{channel1})) is the angular distribution of the resonances. Wolf et al. \cite{wolf} parametrized the angular distribution of the experimental data \cite{ange1} for $p+p \to n+\Delta^{++}$ and assumed the same angular dependence for each isospin channel in the following form, \begin{equation} g_{\rm R}(s,\cos\theta) = a_0(s) \; [ a_1(s) + 3 a_3(s) \cos^2\theta ], \label{angwolf1} \end{equation} with \begin{equation} a_0(s) = \frac{1}{4 \pi ( a_1(s) + a_3(s) ) }. \label{angwolf2} \end{equation} The values of $a_1(s)$ and $a_3(s)$ are given in Table~\ref{table5}. In the high energy region $\sqrt{s} > 2.4 $ GeV, which corresponds to the laboratory energy higher than 1.2 GeV, this angular distribution is assumed to be constant, since there is no experimental data to be fitted in this energy region. However, above the resonance region $E_{{\rm lab}} > 1.2 $~GeV, this is not justified because, for example, the angular distribution of protons from $^{27}$Al$(p,p')$ at 3.17 GeV reaction calculated by Eq.\ (\ref{angwolf1}) deviates from the experimental data (cf. discussions on Fig.\ \ref{enyo1} in the next section). In order to get a better parametrization for high energy part, we assume that the angular dependence is effectively written as a sum of $g_{\rm R}$ and another term $g_{\rm D}$ as \begin{equation} g_{\rm A}(s,\cos\theta) = \frac{1}{2} \, \left[ g_{\rm R}(s,\cos\theta) + g_{\rm D}(s,\cos\theta) \right]. \label{angtot} \end{equation} where \begin{equation} g_{\rm D}(s,\cos\theta) = b_0(s) \; \exp \left[ - 2 \, p^2(s) \, b_1(s) \; ( 1 - \cos\theta ) \right], \label{angdir1} \end{equation} and \begin{eqnarray} b_0(s) & = & \frac{p^2(s) \, b_1(s)} {\pi \, ( \, 1- \exp[ -4 \, p^2(s) \, b_1(s)] \, ) }, \\[3ex] b_1(s) & = & \frac{0.14 \, s^2 \, [3.65 \, ( \sqrt{s} -m_N -m_R) \, ]^6 } {1 + [3.65 \, ( \sqrt{s} -m_N -m_R) \, ]^6 }, \\[3ex] p^2(s) & = & \frac{[s-(m_N-m_R)^2][s-(m_N+m_R)^2]} {4s}. \end{eqnarray} This form of $g_{\rm D}$ is obtained by modifying the Cugnon parametrization \cite{vuu00,cascade01} of the $NN$ elastic angular distribution so as to trace the angular distribution of Eq.~(\ref{angwolf1}) in the resonance region, and approach the elastic-like angular distribution for the higher momentum region. We use this angular distribution for $\Delta$-resonance and $N^*$(1440). For the latter case, we apply this formula by shifting the energy $\sqrt{s}\;$ by the mass difference of the two resonances, i.e., 208 MeV. The energy dependence of this angular distribution is shown in Fig.\ \ref{angf}, where we plot $g_{\rm R}$ (gray bold dashed lines) and $g_{\rm D}$ (solid lines) in (a), while $g_{\rm R}$ (gray bold dashed lines) and $g_{\rm A}$ (solid lines) in (b). In these figures, we symmetrized the elastic-like angular distribution, i.e., $ \frac{1}{2} [ g_{\rm D}(s,\cos\theta) + g_{\rm D}(s,-\cos\theta)] $, in order to compare it with the angular distribution of Eq.\ (\ref{angwolf1}) (gray bold dashed lines). In the next section, we will discuss the dependence of the final results of $(N,xN')$ reactions on this angular distribution of the hadronic resonances. The cross sections for the channel 3 and 5 in Eq.\ (\ref{channel1}) are determined by the law of detailed balance from the cross section of channel 2 and 4 with taking into account the mass dependence of the cross section \cite{wolf2}. For the pion absorption cross section on nucleons (channel 6 and 7 in Eq.\ (\ref{channel1})) we take the maximum cross sections from the particle data table \cite{ptable} and scale them according to the Breit-Wigner formula. For the case of pion absorption on $\Delta$ (channel 8 in Eq.\ (\ref{channel1})), we assume the same cross section as the channel 7 with shifting the energy by the mass difference of the $\Delta$ and nucleon. Apart from the collision term, we take into account the decay of the baryonic resonances during the propagation as \begin{eqnarray} \begin{array}{rlllll} 9. & \Delta & \to & N & + & \pi \\ 10. & N^* & \to & N & + & \pi \\ 11. & N^* & \to & \Delta & + & \pi \\ \end{array} \label{channel2} \end{eqnarray} The decay probability of the resonances is determined by a exponential decay law using their momentum-dependent width Eq.\ (\ref{masdis2}) and their proper time. The decay is assumed to be isotropic in their rest frame. The branching ratio of the channel 10.\ and 11.\ of Eq.\ (\ref{channel2}) is taken from the particle data table \cite{ptable} as $\Gamma_{N^* \to \Delta+\pi} / [\Gamma_{N^*\to \Delta+\pi} + \Gamma_{N^*\to N+\pi}] = 0.4$. The other branching ratios concerning their isospin are determined from the appropriate Clebsch-Gordan coefficients. \subsubsection{Relativistic corrections} Non-covariant framework is another problem of applying the QMD method to the reactions at higher bombarding energies. As explained above, we have already introduced the relativistic form of energy expression in the Hamiltonian Eq.\ (\ref{ham1}) and adopted the relativistic kinematics in the collision term. However a covariant formulation of the interaction term is necessary for a full relativistic description. A Lorentz covariant extension of the QMD, dubbed relativistic quantum molecular dynamics (RQMD), has been proposed by Sorge et al. \cite{rqmd} based on the Poincar\'e-invariant constrained Hamiltonian dynamics. Although the RQMD is a numerically feasible extension of QMD toward a fully covariant approach, it still costs too much computing time to apply the RQMD model to heavy systems. We thus make the following alternative extension of QMD and include the main part of the relativistic dynamical effects in our model. Lehmann et al. \cite{lehmann} have compared the time evolution of the phase space and the particle production obtained by QMD and RQMD, looking for the relativistic effects in heavy-ion collisions in the intermediate energy regime. They found that there is no significant difference between the results of QMD and RQMD in the $\eta$ and $\pi$ meson production cross sections \cite{lehmann} and the proton inclusive spectra \cite{rqmdt}. The difference appeared only in the values of the maximum density \cite{rqmdt} and the transverse flow \cite{rqmdm}. Both are larger in RQMD. Their studies showed that a large part of the difference comes from the Lorentz contraction of the initial phase space distribution in RQMD. If this Lorentz contraction is employed in the normal QMD, however, the transverse flow is overestimated. Thus they concluded that in RQMD this effect is partially counterbalanced by the covariant treatment of the interaction, but there still remains an increased flow compared with the normal QMD calculation. Based on their investigation, we introduce in this paper the Lorentz-scalar quantities into the arguments of the interactions in Eq.\ (\ref{ham1}) as well as the Lorentz contraction of the initial phase space distribution. By these modifications the main part of the relativistic dynamical effects would be approximately included in our QMD. All arguments of the interaction in Eq.\ (\ref{ham1}) are written as a function of the squared spatial distance: \begin{equation} {\bf R}_{ij}^2 = ({\bf R}_i - {\bf R}_j)^2. \end{equation} In the RQMD \cite{lehmann,rqmdt}, these arguments are replaced by the squared transverse four-dimensional distance $-q_{{\rm T}_{ij}}^2$ defined as \begin{equation} -q_{{\rm T}_{ij}}^2 = -q_{ij}^2 + \frac{(q_{ij} \cdot p_{ij})^2} {p_{ij}^2}, \end{equation} where $q_{ij}$ is the four dimensional distance $q_i-q_j$, while $p_{ij}$ denotes the sum of the four momenta of the two particle $p_i+p_j$. In the c.m.s of the particle $i$ and $j$, the squared covariant transverse distance $-q_{{\rm T}_{ij}}^2$ reduces to the usual squared distance. We therefore change the argument in Eq.\ (\ref{ham1}) from the ${\bf R}_{ij}^2$ to the squared distance in the c.m.s. of the two particles $\tilde {\bf R}_{ij}^2$, where the tilde means the quantities defined in the c.m.s of the two particles, \begin{equation} \tilde {\bf R}_{ij}^2 = {\bf R}_{ij}^2 + \gamma_{ij}^2 \left( {\bf R}_{ij} \cdot {\vec \beta}_{ij} \right)^2, \end{equation} with \begin{equation} {\vec \beta}_{ij} = \frac{{\bf P}_i + {\bf P}_j} { E_i + E_j }, \;\;\;\;\;\;\;\; \gamma_{ij} = {1\over \sqrt{1-{\vec \beta}_{ij}^2}}. \end{equation} By this change, all interactions of the Hamiltonian [Eq.\ (\ref{ham1})] depend also on the momentum. The form of the equation of motion [Eq.\ (\ref{newton00})] changes to \begin{eqnarray} \dot{{\bf R}}_i & = & \frac{{\bf P}_i} {\sqrt{m_i^2+{\bf P}_i^2}} + \sum_j D_{ij} \, \frac{\partial \;{\tilde {\bf R}_{ij}^2}} {\partial \,{\bf P}_i}, \label{eom01} \\ \dot{{\bf P}}_i & = & - \sum_j D_{ij} \, \frac{\partial \;{\tilde {\bf R}_{ij}^2}} {\partial \,{\bf R}_i}, \label{eom02} \end{eqnarray} with \begin{eqnarray} D_{ij} & = & - \> \frac{1}{2} \frac{A}{\rho_0} \> \frac{1}{2L} \> \rho_{ij} \nonumber \\ & & - \> \frac{1}{1+\tau} \frac{B}{\rho_0^{\tau}} \> \frac{\tau}{2} \left( < \rho_i>^{\tau-1} + < \rho_j>^{\tau-1} \right) \> \frac{1}{2L} \> \rho_{ij} \nonumber \\ & & + \> \frac{e^2}{2} c_i c_j \left\{ - \frac{1} {|\tilde {\bf R}_{ij}|} \> {\rm erf} \left( \frac{|\tilde {\bf R}_{ij}|}{\sqrt{4L}} \right) + 8 \pi L \> \rho_{ij} \right\} \frac{1}{\tilde {\bf R}_{ij}^2} \nonumber \\ & & - \> \frac{C_{\rm s}}{2\rho_0} \> (1-2|c_i-c_j|) \> \frac{1}{2L} \> \rho_{ij}, \end{eqnarray} and \begin{eqnarray} \frac{\partial \; \tilde {\bf R}_{ij}^2}{\partial \, {\bf R}_i} & = & 2 {\bf R}_{ij} + 2 \gamma_{ij}^2 \left( {\bf R}_{ij} \cdot \vec \beta_{ij} \right) \vec \beta_{ij} \label{dev1} \label{eom11} \\ \frac{\partial \; \tilde {\bf R}_{ij}^2}{\partial \, {\bf P}_i} & = & \frac{2 \gamma_{ij}^2}{E_i + E_j} \left( {\bf R}_{ij} \cdot \vec \beta_{ij} \right) \nonumber \\ & \times & \left\{ {\bf R}_{ij} + \gamma_{ij}^2 \left( {\bf R}_{ij} \cdot \vec \beta_{ij} \right) \left( \vec \beta_{ij} - \frac{ {\bf P}_i}{E_i} \right) \right\} \;, \label{eom12} \end{eqnarray} where $\rho_{ij}$ is defined in Eq.\ (\ref{rhoij}). We also introduce the Lorentz-scalar quantities into the one-body phase-space distribution function Eq.\ (\ref{fi}) as \begin{equation} f_{ij} = 8 \cdot \exp\left[ - {1\over 2L} \> \tilde {\bf R}_{ij}^2 -{2L\over \hbar^2} \> \tilde {\bf P}_{ij}^2 \right] , \end{equation} where $\tilde {\bf P}_{ij}^2$ denotes the squared relative momentum in the c.m.s of the particle $i$ and $j$, which is expressed for two particles with the same mass as \begin{equation} \tilde {\bf P}_{ij}^2 = {\bf P}_{ij}^2 - ( E_i - E_j )^2, \end{equation} with \begin{equation} {\bf P}_{ij} = {\bf P}_i - {\bf P}_j. \end{equation} At the starting point of the QMD calculation, we boost the ground state of projectile (and target as well if c.m.s of the target and projectile is chosen as a reference frame) according to the beam energy. The coordinate ${\bf R}_{iz}^{\rm b}$ and momentum ${\bf P}_{iz}^{\rm b}$ of the nucleon in the beam direction $z$ after the boost are obtained by the Lorentz transformation from those before the boost, ${\bf R}_{iz}$ and ${\bf P}_{iz}$, as \begin{eqnarray} {\bf R}_{iz}^{\rm b} & = & ( {\bf R}_{iz} - {\bf R}_{0z} ) \> / \, \gamma \> + \> {\bf R}_{0z}, \\ {\bf P}_{iz}^{\rm b} & = & \gamma \> ( {\bf P}_{iz} + \beta E_i ), \end{eqnarray} where ${\bf R}_0$ denotes the initial c.m.\ coordinate of the nucleus, while $\beta$ and $\gamma$ are the boosting velocity and its gamma factor, respectively. At this moment, the potential energy of the system and the phase-space distribution function keep the same values as those before the boost due to their Lorentz-scalar properties discussed above. During the propagation of the boosted nuclei, however, those quantities are changing slightly even in the above prescriptions, since the equations of motion Eq.\ (\ref{eom01}) and (\ref{eom02}) are not covariant. But the disturbance due to the non-covariant feature of the equation of motion is negligibly small up to the energy 3 GeV/nucleon. We thus introduce the relativistic corrections discussed above to the non-covariant QMD to save the computing time instead of using the full covariant framework. We should mention here that if one employs the Lorentz contraction for the boosted initial state but does not replace the arguments of the interaction and phase-space distribution by the Lorentz-scalar ones, the potential energy decreases about 80 MeV and the phase-space factor at each nucleon's point changes about 40 \% after the boost in the case of $^{40}{\rm Ca}$ even at 1~GeV/nucleon boosting energy. This means that the boosted ground state obtained by this way may decay spontaneously before it collides with the other nucleus. By our prescription, we are free from this problem. We have checked the above prescription by analyzing the transverse flow, which is sensitive to the treatment of the relativity as discussed in Ref.\ \cite{lehmann}. In Fig.\ \ref{fig-flow} we show the energy dependence of so called directed transverse momentum $<{\bf P}^{{\rm dir}}_x>$, which is a measure for the transverse flow and defined by \begin{equation} <{\bf P}^{{\rm dir}}_x> = {1\over N} \sum_{i=1}^N {\rm sign} \left[ Y_{{\rm CM}}(i) \right] {\bf P}_{ix}, \end{equation} where $Y_{{\rm CM}}(i)$ is the rapidity of the $i$-th baryon in the c.m.s and ${\bf P}_{ix}$ its transverse momentum in the reaction plane. We plot the result of the present QMD simulation (solid line with full boxes) as a differences from that of RQMD \cite{rqmdt} for $^{40}{\rm Ca}$ + $^{40}{\rm Ca}$ reactions at $b$ = 2~fm, for the energy range from 150 MeV/nucleon to 4 GeV/nucleon. In these calculations, we use the same ground states (mentioned below), the same gaussian wave packets and the same interaction (mentioned before). We omitted the Coulomb interaction and two-body collision term for a simplicity. In Fig.\ \ref{fig-flow}, we also plot the other results obtained by the standard QMD (dot-dashed line with open circles) without the initial Lorentz contraction and without the relativistic corrections, and by the standard QMD only with the Lorentz contraction (dashed line with open boxes). As mentioned before, the Lorentz contraction of the initial phase space distribution increases the flow, which is shown by the change from the dot-dashed line to the dashed line. By the full covariant treatment of the interaction, however, this effect is counterbalanced, but still remains an increased flow \cite{lehmann}. As seen in Fig.\ \ref{fig-flow}, our prescription does not deviate so much from the full covariant treatment up to 3 GeV/nucleon. We thus expect that our QMD simulation with the relativistic corrections is very close to the covariant simulation RQMD in this energy regime. At much higher energy, however, our result is decreasing linearly from that of RQMD. This deviation comes from the different treatment of the potential between the RQMD \cite{rqmdt} and our QMD; a Lorentz scalar type in the former, while a time-component of the vector type in the later, respectively. This is understood qualitatively by considering a single particle motion under a fixed external potential $U$. In the Lorentz scalar treatment of the potential $U$, the single particle energy $p_i^0$ is expressed in this simple case as \begin{equation} p_i^0 = \sqrt{p_i^2 + m_i^2 + 2m_i U}. \end{equation} Accordingly the equation of motion is \begin{equation} \dot{{\bf P}}_i = - \frac{m_i}{p^0_i} \frac{\partial U}{\partial {\bf R}_i}. \end{equation} On the other hand in our prescription they are \begin{equation} p_i^0 = \sqrt{p_i^2 + m_i^2 } + U, \end{equation} and \begin{equation} \dot{{\bf P}}_i = - \frac{\partial U}{\partial {\bf R}_i}. \end{equation} In this test calculations, the form of $\frac{\partial U}{\partial {\bf R}_i}$, which is attractive in the beginning, is almost the same in our QMD and in the RQMD. Thus the force of our QMD becomes larger and deviates linearly from that of the RQMD as energy increases. Above 3 GeV/nucleon, therefore, the full covariant prescription is necessary to describe the reactions particularly the nucleus-nucleus collisions. This is out of scope of this paper. Some details of actual numerical calculations should be mentioned here, since all potential terms depend on the momentum by the relativistic corrections. In order to keep a numerical accuracy, we use the second order Runge-Kutta method to integrate the equations of motion. For the energy conservation for the collision term, we assume \begin{equation} E_i + E_j + U_{{\rm pot}} = E_i' + E_j' +U_{{\rm pot}}', \label{econ1} \end{equation} where $E_i$, $E_j$, and $E_i'$, $E_j'$ are the energies of particle $i$ and $j$ before and after the collision, while $U_{{\rm pot}}$ and $U_{{\rm pot}}'$ the potential energy of the system. We determine iteratively the final momenta of the colliding particles so as to satisfy the energy conservation Eq.\ (\ref{econ1}). This prescription is applied to the channels 1 to 5 in Eq.\ (\ref{channel1}), 9 to 11 in Eq.\ (\ref{channel2}). For the pion absorption channels of 6, 7 and 8 in Eq.\ (\ref{channel1}), the energy conservation is written as \begin{equation} E_i + E_j + U_{{\rm pot}} = E_R +U_{{\rm pot}}', \label{econ2} \end{equation} where $E_R$ is the resonance energy. In this case, we determine iteratively the rest mass of the resonance to conserve the energy. \subsubsection{Properties of ``ground state''} An important ingredient of the QMD calculation is how to determine the initial phase space distribution of the projectile and target. For that we cannot use the real ground state (energy minimum state) of the system defined by the Hamiltonian Eq.\ (\ref{ham1}), since the model does not have the Fermionic properties. However, it is necessary to obtain a stable ``ground state''. Also some typical properties of the real ground state should approximately be fulfilled, especially the binding energy and phase space distribution. To get such ``ground state'', we employ the following random packing procedure \cite{f-qmd}; We distribute the centers of position ${\bf R}_i$ of the individual Gaussian wave packet according to a distribution of the Woods-Saxon shape with the radius $R_0 = 1.124 \; A^{1/3} - 0.5 $ fm and the diffuseness parameter $a = 0.2 $ fm. We cut off the Woods-Saxon tail at $R_{{\rm max}} = 1.124 \; A^{1/3} $ fm. In this procedure, we impose minimum distance between the centers of the Gaussians in order to reduce the density fluctuation. We use 1.5 fm for the identical nucleons and 1.0 fm for the other. Now we can calculate the density and potential energy at any point (here, we do not need the relativistic correction discussed in the previous sub-section). Then the center of momentum ${\bf P}_i$ is randomly sampled from the sphere of radius $p_{\rm F}({\bf R}_i)$ which is the Fermi momentum obtained by the local Thomas-Fermi approximation. This sampling is rejected and another value is sampled if the sum of kinetic and potential energy of the particle is positive or the phase space factor $f({\bf R}_j,{\bf P}_j)$ (cf. Eqs.~(\ref{f0}) and (\ref{fi}) ) for the nucleon $j$ which have been previously accepted violates the Pauli principle \cite{f-qmd}. Finally, we check the total binding energy with the simple mass formula \cite{bm1}, i.e., \begin{eqnarray} E_{{\rm bin}} = & - & 15.56 \; A + 17.23 \; A^{2/3} \nonumber \\ & + & 46.57 \; \frac{(N-Z)^2}{2\,A} + \frac{3}{5} \; \frac{Z^2e^2}{1.24\,A^{1/3}}. \label{liquid-e} \end{eqnarray} If the binding energy per nucleon obtained by our sampling lies within $E_{{\rm bin}}\pm 0.5$, we adopt this configuration as a ``ground state''. Thus the ``ground state'' obtained by this procedure always has an appropriate binding energy. However, there is still open phase space below the Fermi surface, since the ``ground state'' is not the energy minimum state of the Fermions. In fact, during the time evolution of the ``ground state'' under the QMD dynamics described in previous sub-section, only 70~\% of the collisions are blocked by the final state Pauli blocking in the two-body collision term. It is allowed at a collision that one nucleon goes down to the lower energy state and the other goes up to the positive energy state. This means that some nucleons could be spontaneously emitted after some time due to the fluctuation of the configuration. To avoid this problem, we assume from a technical point of view that any pair of nucleons originated in the same nucleus does not collide each other until at least one of them experience a collision with a nucleon from the other nucleus. By this assumption the number of the emitted nucleons from the ``ground state'' is reduced to less than about 1~\% of the nucleons up to the time 200 fm/$c$. The density profile of the ``ground state'' obtained here has high density in the center and rather wide surface shape. This is due to the large width of Gaussian $L = 2.0 \; {\rm fm}^2$ used in this paper. On the other hand, the momentum distribution of the ``ground state'' well reproduces the result of Hartree-Fock calculation. In Fig.\ \ref{denmom}, we show (a) the density $\rho(r)$ distribution and (b) the momentum distribution $\rho(p)$ for the ``ground state'' of $^{40}{\rm Ca}$ obtained by QMD simulation (solid lines). The results shown here are averaged quantities over time evolution up to 200 fm/$c$ and over 100 events. The error bars in Fig.\ \ref{denmom} denote the fluctuations in time evolution averaged over 100 events. Although the fluctuation of the one event is much larger, this figure shows that the ground state profile is very stable in time on the average over the events. In the same figure, we also plot the results of Hartree-Fock calculation (dot-dashed lines) and the limit of infinite nuclear matter (dashed lines). The energy spectra of the emitted particles given in the next section, particularly of the subthreshold particle production \cite{niita1} are more sensitive to the momentum distribution than the density profile. This is the reason why we adopted a parametrization which leads to a better momentum distribution at the cost of diffuse density profile. \subsection{Statistical Decay Model} At the end of the dynamical stage of the reactions, the QMD simulation yields many fragments, which are normally in highly excited states. One may think that the decay process of the excited fragments might be described by the QMD dynamics if we can continue the calculation for enough long time. However, we do not follow this method but instead we stop the QMD calculation and switch to the statistical decay model (SDM) at the end of the dynamical stage. There are two reasons for this hybrid model. One is that the time scales of the dynamical and statistical processes are quite different. It is not clever or even practically impossible to continue the reliable QMD calculation more than $10^{-20}$ sec which is necessary to calculate the decay process. Another is a more fundamental reason that the Fermi statistics, which is essential to describe the decay process of the fragments, cannot be traced correctly in the QMD simulation \cite{ohnishi1}. We identify the fragments together with their excitation energies at about 100 $\sim$ 150 fm/$c$ of the QMD simulation. The dependence of the final results on this switching time will be discussed in the next section. Each fragment is recognized by using a minimum distance chain procedure, i.e., two nucleons are considered to be bound in a fragment if the distance between their centroids is smaller than 4 fm. We then calculate the total energy of the fragment in its rest frame and estimate the excitation energy by subtracting the ground state energy given by Eq.\ (\ref{liquid-e}). Though there have been proposed many sophisticated statistical decay codes so far, we use here the simple model of light particle evaporation. We consider only $n$, $p$, $d$, $t$, $^3$He, and $\alpha$ evaporation. The emission probability $P_x$ of these particles $x$ is given with the Fermi gas model as, \begin{equation} P_x = (2J_x +1) \, m_x \, \epsilon \, \sigma_x(\epsilon) \, \rho(E) \, d\epsilon , \label{prob1} \end{equation} where $J_x, m_x,$ and $\epsilon$ are the spin, mass and kinetic energy of the particle $x$, while $\sigma_x(\epsilon)$ and $\rho(E)$ denote the inverse cross section for the absorption of the particle with energy $\epsilon$ and the level density of the residual nucleus with the excitation energy $E$, respectively. We use the following simple form for $\rho(E)$, \begin{equation} \rho(E) = w_0 \, \exp\left( 2 \sqrt{a\, E} \right), \label{level1} \end{equation} with $a = A/8 \; {\rm MeV}^{-1}$ and $w_0$ is a constant. The inverse cross section is assumed to have the form \begin{equation} \sigma_x(\epsilon) = \left\{ \begin{array}{lr} \left(1-U_x/\epsilon\right) \, \pi R^2 & : \epsilon > U_x \\ [1ex] \; 0 & : \epsilon \le U_x \\ \end{array} \right. \label{invers} \end{equation} where $R$ denotes the absorption radius and $U_x$ is a Coulomb barrier for the particle $x$, for which we employ empirical values used in the existing statistical decay code \cite{gemini}. The excitation energy $E$ in Eq.\ (\ref{level1}) is given by \begin{equation} E = E_0 - \epsilon - Q, \label{excite} \end{equation} where $E_0$ denotes the excitation energy of the parent nucleus and $Q$ is the reaction Q-value calculated from the mass formula Eq.\ (\ref{liquid-e}). The total emission probability $R_x$ of the particle $x$ is obtained by integrating the available energy of Eq.\ (\ref{prob1}) as \begin{eqnarray} R_x & = & (2J_x+1) \, m_x \, \nonumber \\[1ex] & & \times \int_{U_x}^{E_0-Q_x} \, \epsilon \, \sigma_x(\epsilon) \, \rho(E_0-Q_x-\epsilon) \, d\epsilon. \label{totalp} \end{eqnarray} This integration can be calculated analytically and the energy spectrum of the emitted particles is given by \begin{equation} N(\epsilon_x) \, d\epsilon_x = \frac{\epsilon_x - U_x}{T_x^2} \, \, \exp\left\{ - \frac{\epsilon_x - U_x}{T_x} \right\} \, \, d\epsilon_x, \label{spec1} \end{equation} with \begin{equation} a\, T_x^2 = E_0 - U_x - Q_x. \label{spec2} \end{equation} In this formulation, we do not consider the $\gamma$ decay nor the angular momentum dependence. The latter is important for the heavy-ion reactions but not so serious for the nucleon-induced reactions considered in this paper. We simulate the whole statistical decay process as a sequential light particle evaporation discussed above by making use of the Monte-Carlo method until no more particle can be emitted. \section{Analysis of the (N,xN') Reactions} In this section, we systematically apply the QMD plus SDM method described in the previous section to $(N,xN')$ (nucleon in, nucleon out) type reactions. In order to get sufficient statistics, we performed the QMD calculations for a large number of events, typically 50000 events, and averaged them to obtain the following results. We first check the dependence on the switching time $t_{{\rm sw}}$ when the QMD calculation is stopped and switched to SDM, which is an ambiguous point of the present model. In Fig.\ \ref{pb15001}, we show a typical neutron energy spectrum at $30^{\circ}$ laboratory angle for the reaction $p(1.5 {\rm GeV}) + ^{208}$Pb. Note that x-axis is plotted in a logarithmic scale to compare the calculated results in detail with the experimental data particularly in a low energy region. The experimental data (full boxes with error bars) are taken from Ref.\ \cite{ishibashi}. The solid histogram denotes the final result of the QMD + SDM calculation. In this case, we switch the QMD calculation to SDM at 100 fm/$c$. In the same figure, we also plot the spectrum of the neutron obtained only by the QMD calculation up to 100 fm/$c$ (dot-dashed histogram) and that coming from the QMD fragments calculated with SDM (dashed histogram), respectively. The former shows a cascade and/or preequilibrium energy spectrum, while the latter an evaporation spectrum. These two components of spectrum are affected by changing the switching time $t_{{\rm sw}}$. However, the total spectrum shape, which is the sum of the two components, stays almost unchanged if an adequately long time is chosen for the switching time $t_{{\rm sw}}$. This is shown in Fig.\ \ref{pb15002}, where we plot results of the total spectra calculated by QMD + SDM with three different switching times, 50 fm/$c$ (dashed line), 100 fm/$c$ (solid line) and 150 fm/$c$ (dot-dashed line). This figure shows that although the latter two lines resemble each other, they deviate definitely from the first line. This indicates that the QMD fragments before 100 fm/$c$ are not in thermal equilibrium and that within a time interval from 100 fm/$c$ to 150 fm/$c$ the decay processes of the excited fragments described by QMD and SDM are nearly equivalent. Although we should keep in mind that these two are not identical at low temperature as the former is always dominated by the classical statistics \cite{ohnishi1}, we can conclude that the final results are not sensitive to the switching time $t_{{\rm sw}}$ as long as it is chosen after the time when the thermal equilibrium is achieved and before the time the temperature of the fragments become low and classical statistics breaks down seriously. The similar conclusion has been obtained in Ref.\ \cite{maru92a}, which indicates that the minimum switching time to get stable results depends on the size of system and the incident energy. For the case of nucleon induced reactions, we found that 100 fm/$c$ is enough to get stable results and we use this value for all systems in the present study. Next check is the detailed examination of the inelastic channels in the two-body collision term. For this purpose we compare our results with the experimental data at high incident energy of proton on the light-mass target, which directly reflects the elementary processes included in the model. In Fig.\ \ref{enyo1}, we plot the invariant cross section of the proton (left-hand-side) and negative pion (right-hand-side) emission for the reaction $p \,(3.17 {\rm GeV}) +^{27}$Al. The experimental data (full boxes with error bars) are taken from Ref.\ \cite{enyos} and the results of QMD + SDM are denoted by solid histograms. In the same figure, we plot results of QMD + SDM with the different choice of the angular distribution of the resonances. The dashed histograms are the results obtained with only the resonance-like angular distribution of Eq.\ (\ref{angwolf1}), while the dot-dashed histograms are those with the direct-like angular distribution of Eq.\ (\ref{angdir1}), respectively. This figure shows that the average of the two components of the angular distribution of Eq.\ (\ref{angtot}) well fits the experimental proton spectra. On the other hand, the pion spectra are rather insensitive to the angular distribution of the resonances. Instead, their spectra are dominantly determined by the mass distribution of the resonances of Eq.\ (\ref{masdis1}). Although the authors of Ref.\ \cite{enyos} analyzed these data by making use of the two-moving-source model, our QMD + SDM can reproduce excellently the proton and pion spectra at the same time without any special assumption. In Figs.\ \ref{fe113}-\ref{pb3000}, we compare the neutron energy spectra obtained by the QMD + SDM calculations with the experimental data \cite{ishibashi,Meier} for Fe and Pb targets at the proton energy from 113 MeV up to 3 GeV. In the fields of application of accelerators, such as spallation neutron sources, accelerator-based transmutation systems, and shielding of cosmic ray in space activity, the production of slow neutrons plays an important role. That is the reason we chose these data \cite{ishibashi,Meier} to compare with, since the neutron spectra from 1 MeV up to the beam energy are available. For efficient comparisons of the calculations and the data, both for low energy and high energy regions, we plot the same results in two figures with the x-axis in a logarithmic scale (left-hand-side) and in a linear scale (right-hand-side). One can see the detail of the thermal and preequilibrium neutron spectra in the former figure, while the direct-like components of the spectra in the latter figure, respectively. These figures indicate that over the broad range of the incident energies from 100 MeV to 3 GeV, independently of the targets, and of all angles of the outgoing neutrons, our results of the neutron energy spectra agree well with the data from 1 MeV up to the beam energy. Though one may notice some disagreement at the high energy part of the most forward angle, which will be discussed later, the overall agreement is satisfactory. Particularly a remarkable agreement of the present calculations with the low energy neutron data below several tens MeV confirms that the QMD gives proper excitation spectra of the excited fragments from which the statistical neutron emission takes place. With a suitable chosen fixed set of parameters, our QMD plus SDM model is able to reproduce quantitatively the overall neutron spectra for the broad range of the incident energies and target masses. At 113 MeV, we additionally compare our results with the prediction of the intranuclear cascade plus light particle evaporation model (NUCLEUS) \cite{nucleus}. This model is essentially the intranuclear reaction part of NMTC \cite{nmtc} and HETC \cite{hetc} codes. Calculations with NUCLEUS yield almost the same results as ours in the forward angles but give lower values in the backward angle, which is denoted by the dashed histogram at 150 degree in Fig.\ \ref{fe113}. In this energy regime it has been reported \cite{blann1,blann2} that the semiclassical preequilibrium models based on an intranuclear cascade model also fail to reproduce the angular distributions. We found from the detailed comparison of the calculations that the underestimation of the backward angle in the above models is due to the insufficient treatment of the soft interaction of a nucleon with all the rest of the nucleons in the nucleus, which is naturally included in the QMD formalism. One may think that this explanation is in contradiction with Ref.\ \cite{peilert,blann3} where they attribute the failure to the insufficient contributions from second and higher order collisions. To resolve this problem, we have checked that the NUCLEUS has almost the same prescription of hard nucleon-nucleon interaction and has almost the same momentum distribution in the ground state as QMD. Difference is that our QMD part includes the soft nucleon-nucleon interaction but NUCLEUS does not. This soft interaction diffracts the nucleon. As a result, the yields of the backward angle increase and by the same reason the number of multiple hard collisions also increases. On the other hand, a multistep model of the Feshbach-Kerman-Koonin (FKK) \cite{fkk} has been also applied to the energy regime around 100 MeV \cite{blann3}. Although the FKK approach successively reproduced the angular distribution, the overall absolute values of their results are very sensitive to the strength parameter of the residual interaction, which is adjusted to fit the experimental data. The strength so determined depends on the incident particle, target nucleus and incident energy. In QMD, on the contrary, the parameters of the soft nucleon-nucleon interaction in Eq.~(\ref{ham1}) are taken common to all reactions and determined from the nuclear saturation condition. In addition, the final results are not so sensitive to them. The first analysis of $(p,xn)$ reactions by the QMD approach in the energy regime up to 800 MeV has been done by Peilert et al. \cite{peilert}. The neutron spectra of their results are very similar to those of the present work above several tens MeV. Their analysis, however, cannot predict a whole spectra of neutron, since the contribution of the statistical decay from the excited fragments produced in the QMD calculation was not considered in their work. Though the present results show overall agreement with the data for the very broad energy regime, one can see a systematic deviation from the data in the high energy part of the neutron spectra at the most forward angle at incident energy from 113 MeV up to 800 MeV (see the right-hand-side of Fig.\ \ref{fe113}-\ref{pb800}). We suppose that the soft nucleon-nucleon interaction is responsible for this deviation. One possibility is a momentum dependent interaction that is not included in the present QMD, by which the nucleon could be affected coherently by the surrounding nucleons when its momentum is drastically changed by the hard nucleon-nucleon scattering. For the higher incident energies (see Fig. \ref{pb1500} and \ref{pb3000}), this deviation disappears. In those cases, we have checked that the neutrons in the high energy part of the forward angle emerge after at least once experiencing the resonances of nucleon, and that the effects of the soft interaction is relatively small. The analysis by the QMD including the momentum dependent interaction will be reported in the forthcoming paper. \section{Summary and Conclusion} We have proposed the quantum molecular dynamics (QMD) incorporated with the statistical decay model (SDM) aiming to describe various nuclear reactions in a unified way, and applied this model to the $(N,xN')$ reactions. We have checked and found that the final results do not depend on the switching time when the QMD simulation is stopped and switched to the SDM calculation as long as the switching time is chosen between 100 fm/$c$ and 150 fm/$c$ for the nucleon induced reactions. Therefore there left little ambiguity with respect to the switching of two different kinds of models to describe whole process of the reactions in a unified way. In order to describe the reactions at high incident energies up to 3 GeV, we have taken into account two baryonic resonances, the $\Delta$(1232) and the $N^*$(1440) as well as the pions in the QMD model. The elementary cross sections related to these resonances and pions are basically taken from the experimental data. The angular distributions of the resonances, which information is very poor in the experimental data, have been fixed to fit the $^{27}$Al$(p,p')$ data \cite{enyos} at 3.17 GeV. It should be noted that the energy spectra of the nucleons from the $(N,xN')$ reactions on the small target are suitable quantities to obtain the detailed information of the angular distribution of the resonances, while the pion spectra are useful to extract the information of the mass distribution of the resonances. In addition to the relativistic kinematics and approximately covariant prescription of the collision term, we have introduced the Lorentz-scalar quantities to the arguments of the interactions and to the phase space factor. By this relativistic corrections together with the Lorentz contraction of the initial phase space distribution, the main part of the relativistic dynamical effects is approximately described in our QMD for the energy regime up to 3 GeV/nucleon. Validity of this model has been confirmed by the analysis of the transverse flow for the heavy-ion collisions in comparison with the results obtained by the covariant version of quantum molecular dynamics (RQMD). We have applied systematically QMD + SDM to the $(N,xN')$ reactions for a broad range of incident energies from 100 MeV to 3 GeV and of target masses. The present model reproduced the overall feature of the outgoing neutrons quite well without assuming any reaction mechanism, and without changing a parameter set. Although there are a lot of parameters in the model which have not been investigated extensively in this paper, the final neutron spectra analyzed here do not depend so much on them, for example, the equation of state (choice of the interaction), width of the gaussian wave packets, and the detail of the statistical decay process. The main ingredients of the model, which produce the present results of the neutron spectra down to the energy of several MeV, are the parametrization of the elastic and inelastic elementary cross section and the many-body dynamics itself, which have been discussed both in detail in this paper. We thus conclude that the present QMD + SDM scheme gives a unified picture of the major three reaction mechanisms of $(N,xN')$ reactions; i.e. compound, pre-equilibrium and spallation processes. Finally, we should mention that the present model is ready to be applied directly to the heavy-ion reactions in its original form. A study of heavy-ion reaction using this model is now under consideration. \acknowledgements The authors are grateful to Profs. K. Ishibashi and M. M. Meier for supplying us with the experimental data prior to the publication.
\section{Introduction} QCD in two space-time dimensions ($QCD_2$) is believed to be a useful laboratory to examine ideas about the hadronic physics of the real world. One such an idea, the expansion in large number of colors $N_c$, was applied in the pioneering work of 't Hooft \cite{thooft}. A mesonic spectrum of a ``Regge trajectory" type was discovered in that work for $QCD_2$ with Dirac fermions in the fundamental representation. The large $N_c$ limit is taken while keeping $e^2N_c$ fixed, where $e$ is the gauge coupling. The ``orthogonal" approach, the strong coupling limit, was also found to be a fruitful technique when combined with bosonizing the system. In that framework the low energy effective action of the theory can be derived exactly. Quantization of soliton solutions of that effective theory led to the low lying semi-classical baryonic spectrum\cite{DFS}. The large $N_c$ approach combined with a light-front quantization was applied to the study of $QCD_2\ $ with Majorana fermions in the adjoint representation. Unlike the case of fundamental fermions, for adjoint fermions pair creation is not suppressed in the large $N_c$ limit. The bosonic spectrum includes an infinite number of approximately linear Regge trajectories which are associated with an exponential growing density of states at high-energy\cite{kutasov,gyan}. Recently, a ``universality" behavior of $QCD_2\ $ was found in the sense that the physics of massive states depends only on the gauge group and the afine Lie algebra level but not on the representation of the group\cite{kutasov2}. In the present paper we combine the bosonization technique with that of a large $N$ expansion and a light-front quantization in the analysis of the massive mesonic spectrum of several $QCD_2\ $ models. The massless sector which was discussed in ref.\cite{kutasov2} is not addressed in the present work. The models include massless fermions in the adjoint representation and multi flavor fundamental representations. In the former case we expand in $N_c-$ the number of colors whereas in the latter case we consider large $N_f-$ number of flavors . In case of massless fermions the bosonization formalism is convenient since it separates the color, flavor and baryon number degrees of freedom\cite{FSrep}. Moreover the generalization from fermions coupled to YM fields to an arbitrary gauged afine Lie algebra system is natural in the bosonized picture. The basic difference between our approach and the one taken in refs. \cite{kutasov,gyan} is the use of current quanta rather then those of quarks in constructing the mass operator $M^2$ and thus also the wave equation and the Hilbert space. Our results are in agreement with the results of refs.\cite{kutasov,gyan} for the case of adjoint fermions. In the large $N_f$ it is shown that the exact massive spectrum is a single particle with $M^2={{e^2 N_f} \over \pi}$. This phenomenon is explained by the fact that this limit can be viewed as an ``abelianization" of the model. The paper is organized as follows. In section 2 we present the models. The actions are written down in their bosonized versions. We then derive explicit expressions for the momentum operators and the mass operator. The afine Lie algebra currents are expanded in terms of annihilation and creation operators with which a Fock space of physical states is built. In section 3 the model with adjoint fermions is analyzed in the large $N_c$ limit. We introduce wave functions and write down an eigenvalue equation which generalizes 't Hooft equation\cite{thooft}. The massive mesonic spectrum was then deduced. In section 4 we find that the exact spectrum of multi-flavor QCD with fundamental fermions in the regime of $N_f \gg N_c$ include only one single mesonic state. We discuss the relation of this spectrum to that of multi-flavor QED. The case of large level $WZW$ models is also considered. Some conclusions and certain open problems are discussed in section 5. \section{The Models} Consider two dimensional $SU(N_c)$ Yang-Mills gauge fields coupled to (i) $N_f$ massless Dirac fermions in the fundamental representation, or (ii) massless Majorana fermions in the adjoint representation. These theories are described by the following classical Lagrangian: \beq {\cal L}=-{1\over {4e^2}}Tr[ F_{\mu\nu}^2+i\bar \Psi\not{\hbox{\kern-4pt $D$}}\Psi] \eeq where $F_{\mu\nu}=\partial _\mu A_\nu - \partial _\nu A_\mu + i[A_\mu ,A_\nu]$ and the trace is over the color and flavor indices. For the case (i) $\Psi$ has the following group structure $\Psi_{ia}$ where $i=1,...,N_c$ and $a=1,...,N_f$, $D_\mu=\partial_\mu -iA_\mu$ whereas for the case (ii) $\Psi\equiv \Psi^i_j$ and $D_\mu=\partial_\mu -i[A_\mu,\ ]$. In both cases $\Psi$ is a two spinor parametrized as follows $\Psi = \left( \begin{array}{c} \bar \psi \\ \psi \end{array} \right) $. It is useful to handle these models in the framework of the light-front quantization, namely, to use light-cone space-time coordinates and to choose the chiral gauge $A_-=0$. In this scheme the quantum Lagrangian takes the form \beq {\cal L}=-{1\over {2e^2}} {(\partial _-A_+)}^2+i\psi ^\dagger \partial_+\psi +i\bar \psi ^\dagger \partial_-\bar \psi + A_+J^+ \eeq where color and flavor indices were omitted and $J^+$ denotes the $+$ component of the color current $J^+ \equiv \psi ^\dagger \psi$. By choosing $x^+$ to be the `time' coordinate it is clear that $A_+$ and $\bar \psi$ are non-dynamical degrees of freedom. In fact, $\bar \psi$ are decoupled from the other fields so in order to extract the physics of the dynamical degree of freedom one has to functionally integrate over $A_+$. The result of this integration is the following simplified Lagrangian \beq {\cal L}={\cal L}_0+{\cal L}_I=i\psi ^\dagger \partial_+\psi+i\bar\psi ^\dagger \partial_-\bar\psi -e^2J^+{1\over {\partial _- ^2}}J^+\label{SimAc} \eeq Since our basic idea is to solve the system in terms of the ``quanta" of the colored currents, it is natural to introduce bosonization descriptions of the various models. (i) The bosonized action of colored flavored Dirac fermions in the fundamental representation is expressed in terms of a WZW\cite{witten} action of a group element $u\in U(N_c\timesN_f)$ with an additional mass term that couples the color, flavor and baryon number sectors\cite{FS2}. In the massless case when the latter term is missing the action takes the following form \beq S^{fun}_0=S^{WZW}_{(N_f)}(g) + S^{WZW}_{(N_c)}(h)+ {1 \over 2} \int d^2x \partial _\mu \phi \partial ^\mu \phi\label{BLfc} \eeq where $g\in SU(N_c),\ h\in SU(N_f)$ and $e^{i\sqrt{4\pi\over N_cN_f}\phi}\in U_B(1)$ with $U_B(1)$ denoting the baryon number symmetry and \ber \lefteqn{S^{WZW}_{(k)}(g)={k\over{8\pi}}\int _\Sigma d^2x \ tr(\partial _\mu g\partial ^\mu g^{-1}) + } \\ && {k\over{12\pi}}\int _B d^3y \epsilon ^{ijk} \ tr(g^{-1}\partial _i g) (g^{-1}\partial _j g)(g^{-1}\partial _k g), \eer (ii) The current structure of free Majorana fermions in the adjoint representation can be recast in terms of a WZW action of level $k=N_c$, namely, $ S^{adj}_0=S^{WZW}_{(N_c)}(g)$ where now $g$ is in the adjoint representation of $SU(N_c)$, so that it carries a conformal dimensions of ${1\over 2}$\cite{usss}. Multi-flavor adjoint fermions can be described as $S^{WZW}_{N_f}(g) + S^{WZW}_{N_c^2-1}(h)$ where $g\in SO(N_c^2-1)$ and $h\in SO(N_f)$. In the present work we discuss only gauging of $SU(N_c)$ WZW so the latter model would not be considered. Substituting now $S^{fun}_0$ or $S^{adj}_0$ for $S_0$ the action that corresponds to \eqref{SimAc} is given by \beq S= S_0-{e^2 \over 2}\int d^2x J^+{1\over{\partial _- ^2}}J^+ \eeq where the current $J^+$ now reads $J^+=i{k\over{2\pi}}g\partial _- g^\dagger$, where the level $k=N_f$ and $k=N_c$ for the multi-flavor fundamental and adjoint cases respectively. The light-front quantization scheme is very convenient because the corresponding momenta generators $P^+$ and $P^-$ can be expressed only in terms of $J^+$. We would like to emphasize that this holds only for the massless case. Using the Sugawara construction, the contribution of the colored currents to the momentum operator, $P^+$, takes the following simple form: \beq P^+={1\over{N+k}}\int dx^- :J^i_j(x^-)J^j_i(x^-): \eeq where $J \equiv \sqrt \pi J^+$, $N_c$ in the denominator is the second Casimir operator of the adjoint representation and the level $k$ takes the values mentioned above. Note that for future purposes we have added the color indices $i,j=1 \ldots N_c$ to the currents. In the absence of the interaction with the gauge fields the second momentum operator, $P^-$ vanishes. For the various $QCD_2$ models it is given by \beq P^-=-{e^2 \over {2\pi}}\int dx^-:J^i_j(x^-){1\over{\partial _- ^2}}J^j_i(x^-): \eeq In order to find the massive spectrum of the model we should diagonalize the mass operator $M^2=2P^+P^-$. Our task is therefore to solve the eigenvalue equation \beq 2P^+P^-\state{\psi} = M^2\state{\psi}. \eeq We write $P^+$ and $P^-$ in term of the Fourier transform of $J(x^-)$ defined by $J(p^+)=\int {dx^- \over {\sqrt{2\pi}}} e^{-ip^+x^-} J(x^-)$. Normal ordering in the expression of $P^+$ and $P^-$ are naturally with respect to $p$, where $p<0$ denotes a creation operator. To simplify the notation we write from here on $p$ instead of $p^+$. In terms of these variables the momenta generators are \bea &P^+={2\over{N+k}}\int ^\infty _0dp J^i_j(-p)J^j_i(p) \nonumber \\ &P^-={e^2 \over \pi}\int ^\infty _0dp {1\over{p^2}}J^i_j(-p)J^j_i(p) \label{P} \eea Recall that the light-cone currents $J^i_j(p)$ obey a level $k$, $SU(N_c)$ affine Lie algebra \beq [J^k_i(p),J^n_l(p')]={1\over 2}kp(\delta ^n_i\delta ^k_l - {1\over N}\delta ^k_i \delta ^n_l)\delta (p+p')+ \half (J^n_i(p+p')\delta ^k_l - J^k_l(p+p')\delta ^n_i) \label{Kac} \eeq We can now construct the Hilbert space. The vacuum $\state{0,R}$ is defined by the annihilation property: \beq \forall p>0,\ J(p)\state{0,R}=0 \eeq Where R is an ``allowed" representations depending on the level\cite{gepner}. Thus, a typical state in Hilbert space is \\ $Tr\ J(-p_1)\ldots J(-p_n)\state{0,R}$. Diagonalizing $M^2$, is in general, a complicated task, hence we will examine in detail some special cases of the theory. \section{Large $N_c$ Adjoint Fermions} The first case we analyze is $QCD_2$ with fermions in the adjoint representation. This model was investigated in the past \cite{kutasov} \cite{gyan} and recently it was shown \cite{kutasov2} that its massive spectrum is the same as the model of $N_f=N_c$ fundamental quarks. Since an exact solution of the model is beyond our reach we employ a large $N=N_c$ approximation where some simplifications occur. As stated above the $N_c=N$ is also the corresponding level of the model. First, we write down the general symmetric singlet states: \ber \state{\psi} = \sum ^\infty _{n=2} {1\over {N^n}} \int ^P _0 \ldots \int ^P _0 dp_1 \ldots dp_n \delta (P-\sum ^n _{i=1} p_i) \phi _n (p_1,\ldots ,p_n)\times \\ \times {1\over n} \sum_\sigma J^{j_1}_{j_2}(-p_{\sigma (1)}) J^{j_2}_{j_3}(-p_{\sigma(2)}) \ldots J^{j_n}_{j_1}(-p_{\sigma (n)}) \state{0} \eer \ber \state{\psi} '= \sum ^\infty _{n=2} {1\over {N^n}} \int ^P _0 \ldots \int ^P _0 dp_1 \ldots dp_n \delta (P-\sum ^n _{i=1} p_i) \chi _n (p_1,\ldots ,p_n)\times \\ \times {1\over n} \sum_\sigma J^{j_1}_{j_2}(-p_{\sigma (1)}) J^{j_2}_{j_3}(-p_{\sigma(2)}) \ldots J^{j_n}_{j_{n+1}}(-p_{\sigma (n)}) \state{0} ^{j_{n+1}}_{j_1} \eer Here $\phi _n,\chi _n$ are the Schr\"{o}dinger wave functions, and the summation over $\sigma$ is, as explained below, over the cyclic group $Z_n$. The difference between the two states is that in $\state{\psi}$ the currents act on the identity vacuum, whereas in $\state{\psi} '$ the currents act on the adjoint vacuum. The latter is a state that transforms in the adjoint representation of $SU(N)$ and obeys $J(p) \state{0} ^i_j=0$ for all $p>0$. The two states form two distinct equivalence classes, in a sense that one state can not be related to another one by acting with an operator which is built from a finite number of current operators. In principle, other vacuum representations are allowed, but is was proven \cite{kutasov2} that in the large $N$ limit it is sufficient to consider only the identity vacuum and the adjoint vacuum. Other representations lead to multi particle states which are suppressed in the large $N$ limit. Note that $M^2 \state{0}=0$, but $M^2 \state{0} ^i_j = m_0^2 \state{0} ^i_j \ne 0$. The reason is that the zero mode of the current $J^i_j(0)$ does not annihilate the adjoint vacuum and actually $J^k_l(0)J^l_k(0) \state{0} ^i_j = N \state{0} ^i_j$\cite{gepner} .The factor N is the second Casimir of the adjoint representation. Thus $M^2 \state{0} ^i_j = 2P^+P^- \state{0} ^i_j = 2\times {1 \over {2N}}\times N \times {e^2 \over {2\pi}} \times N \state{0} ^i_j = {e^2 N \over {2\pi}} \state{0}^i_j $. This non-zero value of $m_0^2$ will lead to mass splitting between eigenvalues of $\state{\psi}$ and $\state{\psi} '$. The summation of $\sigma$ over all elements of the cyclic group $Z_n$ was introduce to achieve the following symmetry property of the wave functions \bea \phi _n(p_1,p_2,\ldots ,p_{n-1},p_n)=\phi _n(p_n,p_1,p_2,\ldots ,p_{n-1}) \nonumber \\ \chi _n(p_1,p_2,\ldots ,p_{n-1},p_n)=\chi _n(p_n,p_1,p_2,\ldots ,p_{n-1}) \label{bc1} \eea Another property of $\phi _n,\chi_n$ is their boundary condition \bea \phi _n(0,p_2,\ldots ,p_n)=0 \nonumber \\ \chi _n(0,p_2,\ldots ,p_n)=0 \label{bc2} \eea which is a consequence of hermiticity of the operators\cite{thooft}. A general state whether of the type $\state{\psi}$ or $\state{\psi} '$ is an eigenstate of $P^+$ with eigenvalue $P$ \beq P^+\state{\psi} =P\state{\psi} \eeq \beq P^+\state{\psi} '=P\state{\psi} ' \eeq due to the following commutation relation \beq [P^+,J^{j_1}_{j_2}(-p)]=pJ^{j_1}_{j_2}(-p). \eeq In the more familiar CFT terminology this relation translates into $[L_0, J_n]= -nJ$. The calculation of the commutator $[P^-,J^{j_1}_{j_2}(-p)]$ is more complicated. In the procedure of evaluating the spectrum we find the eigen wave-function by solving an integral equation that generally mixes $\phi_n$ (or $\chi_n$) with different $n$'s, namely, it mixes the number of currents. Since the general equation is highly non-trivial we will use an approximation in which we take into account only singular terms and terms with the same number of current operators. Those terms have the dominant contribution to the wave equation\cite{kutasov}. A detailed calculation of the commutators which includes all the terms is written in the Appendix. Here we write only the most significant term of the commutator \bea \lefteqn {[P^-,J^{j_1}_{j_2}(-p_1)]= } \nonumber \\ && {e^2\over \pi}\half \int ^\infty _0 dp({1\over {{(p\!+\!p_1)}^2}} - {1\over p^2}) (J^i_{j_2}(-p\!-\!p_1)J^{j_1}_i(p)-J^{j_1}_i(-p\!-\!p_1)J^i_{j_2}(p)) \nonumber \\ && + other\ terms \eea The above term includes an annihilation operator $J(p)$, which has non-zero contribution when it commutes with another creation operator, say $J(-p')$. A non-negligible commutator occurs only with a `neighbor' of $J^{j_1}_{j_2}(-p_1)$, namely, a commutator with a current which carries the same group indices. Other commutators are suppressed by factors of ${1\over N}$. A tedious though straightforward calculation yields \bea &\half \int ^\infty _0 dp({1\over {{(p\!+\!p_1)}^2}} - {1\over p^2})[(J^i_{j_2}(-p\!-\!p_1)J^{j_1}_i(p)-J^{j_1}_i(-p\!-\!p_1)J^i_{j_2}(p), J^{j_2}_{j_3}(-p_2)]= \nonumber \\ &{N\over 2}\times \int ^{p_2}_0 dp({1\over {{(p_1\!+\!p_2\!-\!p)}^2}}-{1\over {{(p_2\!-\!p)}^2}}) J^{j_1}_i(-p_1\!-\!p_2\!+\!p)J^i_{j_3}(-p) + other\ terms \eea were `other terms' are either terms that are suppressed in large $N$, or terms that change the number of currents. We comment on the validity of dropping the latter terms in section 5. The result of the last two commutators is that if one starts with two currents $J(-p_1)$ and $J(-p_2)$ he ends with two other currents which are multiplied by singular denominators. This result leads to the following eigenvalue equation \bea M^2\phi _n(p_1,\ldots ,p_n) = -{e^2N \over \pi}\int dy {1\over {(p_1-y)}^2} \phi _n(y,p_1\!+\!p_2\!-\!y,p_3,\ldots ,p_n) + \nonumber \\ cyclic\ permutations \ \ \ \ \ \ \label{thoofteq} \eea This is a generalization of \begin{em} 't Hooft's equation. \end{em} Obviously, the same equation holds also for $\chi _n$. Equation \eqref{thoofteq} can be solved analytically with the boundary conditions \eqref{bc1} and \eqref{bc2}. The simplest solution is for $\phi _2$: \ber \phi _2(x) = \sin (\pi kx) & k\in 2Z+1 \nonumber \\ M^2_k = e^2N\pi k \eer The general solution for $\phi_{2n}$ is quite involved but the eigenvalues are rather simple: \beq M^2_{k_1,k_2,\ldots ,k_n}= e^2N\pi (k_1\!+\!k_2\!+\! \ldots \!+\! k_n) \ \ \ \ k_1,k_2,\ldots ,k_n \in 2Z+1 \eeq The result for the $\state{\psi} '$ sector is similar, with the small $m_0^2$ difference: \beq M^2_{k_1,k_2,\ldots ,k_n}= e^2N\pi (k_1\!+\!k_2\!+\! \ldots \!+\! k_n) + m_0^2 \ \ \ \ k_1,k_2,\ldots ,k_n \in 2Z+1 \eeq This expression of the eigenvalues indicates an exponential growth of the number of states, in accordance with previous results \cite{kutasov}. There is, however, a slight difference between our results and those of ref.\cite{kutasov}. We found that the values of the integers must be odd whereas in Kutasov's paper, they are even. The source of this difference is the symmetric boundary conditions that we have used \eqref{bc1}. The spectrum contains two blocks. The identity-block has integer dimension and hence can be interpreted as the boson-block,whereas the adjoint-block has half integer dimension and thus will be referred as the fermion-block. We have seen that the two blocks have similar structure since the two sectors obey the same wave function equation. Furthermore, mixture between the two sectors is avoided due to the fact that the hamiltonian creates and destroys even numbers of ``quarks". The``quark" content of the spectrum is determined by the relation between the currents and the quarks. This relation is given by \beq J^i_j(-p)=\int ^\infty _{-\infty } dq \psi ^i_k(q)\psi ^k_j(-p-q) \eeq which is the Fourier transform of $J(x)=\psi (x)\psi (x)$, and hence the "boson -block" can be written as \ber \lefteqn{\sum_\sigma J^{j_1}_{j_2}(-p_{\sigma (1)}) J^{j_2}_{j_3}(-p_{\sigma(2)}) \ldots J^{j_n}_{j_1}(-p_{\sigma (n)}) \state{0} = } \\ && \sum_\sigma \int ^\infty _{-\infty} dq_1 \psi ^{j_1}_{i_1}(q_1) \psi ^{i_1}_{j_2}(-p_{\sigma(1)}-q_1) \int ^\infty _{-\infty} dq_2 \psi ^{j_2}_{i_2}(q_2) \psi ^{i_2}_{j_3}(-p_{\sigma(2)}-q_2) \\ && \ldots \int ^\infty _{-\infty} dq_n \psi ^{j_n}_{i_n}(q_n) \psi ^{i_n}_{j_1}(-p_{\sigma(n)}-q_n) \state{0} \eer and can be thought of as a mixture of $2n,2n-2,\ldots ,2$ quarks. In a similar way, a state in the fermion-block is a mixture of $2n+1,2n-1,\ldots ,3$ quarks. \section{Large $N_f$ QCD} Massless $QCD_2$ with $N_f$ flavors of quarks in the fundamental representation is described by the Lagrangian of \eqref{BLfc}. Setting aside the flavor and baryon number sectors, the left over system is that of a $k=N_f$ level $SU(N_c)$ WZW action with an additional non-local interaction term. We are thus led to analyze the spectrum of the model with level equal to $N_f$. In practical terms the latter means substituting $k$ by $N_f$ in the expression for $P^+$ eqn. \eqref{P} and in the Affine Lie algebra eqn. \eqref{Kac}. The idea is to invoke a large $N_f$ approximation, namely, to consider models in which $k=N_f \gg N_c$. Models with number of colors and flavors which fall into this regime, are significantly simpler than models that don't obey this inequality. The basic reason for that is the simplification of the algebra eqn. \eqref{Kac}. The commutator of $P^-$ with $J$ in the large $N_f$ limit takes the form \beq [P^-,J^i_j(-p)]={e^2 N_f\over {2\pi p}} J^i_j(-p) + e^2 JJ\ term \eeq which means that \beq M^2\ {J^i_j(-p) \over {N_f^{1\over 2}}} \state{0,R} = {e^2N_f \over \pi} \ {J^i_j(-p) \over {N_f^{1\over 2}}} \state{0,R} + e^2 N_f^{1\over 2} {JJ \over N_f} \state {0,R}\label{Slnf} \eeq Upon neglecting the second term which is suppressed by a factor of ${1\over {\sqrtN_f}}$ we get the following solution to the eigenvalue problem \beq M^2\ J^i_j(-p) \state{0}^{ja}_{ib} = {e^2N_f \over \pi} \ J^i_j(-p) \state{0}^{ja}_{ib}. \eeq Notice that the current operator acts on the adjoint vacuum and not on the identity vacuum. The reason for that is obviously the requirement that states have to be color singlets. $\state{0}^{ja}_{ib}$ stands for $\state{0}^{j}_{i}\otimes \state{0}^{a}_{b}$, namely, the tensor product of the color adjoint vacuum and its flavor counterpart. Recall that following eqn. \eqref{BLfc} the vacuum state is an outer product of the vacua of the three independent Hilbert-spaces of the color, flavor and baryon number sectors. This ``decoupling" of these spaces of states is an artifact of the massless limit of the bosonized picture. It is further discussed in section 5. The above state is the only one-particle state of the theory. All other massive state are multi particle state which are built from this 'meson'. The content of massless multi-flavor $QCD_2$ in the large $N_f$ limit is thus very simple and is in fact closely related to multi-flavor massless QED\cite{rabin}. In the latter case model the spectrum of single particles contains the following single state \beq M^2\ J(-p) \state{0} = {e^2 N_f \over \pi} \ J(-p) \state{0} \eeq The reason of this similarity is very clear. Neglecting the $O(N_f^{-{1\over 2}})$ term in eqn.\eqref{Slnf} corresponds to dropping the structure constant term in the Affine-Lie algebra. In other words, in this limit we perform an Abelianization of the theory. However, one may identify a difference in the substructure of the two mesons. The Abelian meson is built up from two quarks since current quanta is made out of two quarks, whereas the non Abelian meson is a more complex object. The difference can be seen by writing the state explicitly in terms of quarks operators. The (traceless) current is written as: \beq J^i_j(-p) = \int ^\infty _{-\infty} dq \psi ^{\dagger i}_c(q)\psi ^c_j(-p-q) - {{\delta ^i _j} \over N_c} \int ^\infty _{-\infty} dq \psi ^{\dagger k}_c(q)\psi ^c_k(-p-q) \eeq The flavored adjoint vacuum is written as: \beq \state{0} ^{ja}_{ib} = \psi ^{\dagger j}_b(0) \psi ^a_i(0) \state{0} \eeq It is a state of two quarks with zero (light-cone) momentum acting on the identity vacuum. Thus the massive meson is \ber & J^i_j(-p) \state{0}^{ja}_{ib} = (\int ^\infty _{-\infty} dq [\psi ^{\dagger i}_c(q)\psi ^c_j(-p-q) - {{\delta ^i _j} \over N} \psi ^{\dagger k}_c(q)\psi ^c_k(-p-q)])\times \psi ^{\dagger j}_b(0) \psi ^a_i(0) \state{0}\nonumber \\ & = \int ^p_0 dq \psi ^{\dagger i}_c(-q) \psi ^c_j(-p+q) \psi ^{\dagger j}_b(0) \psi^a_i(0) \state{0} - {1\over N} \int ^p_0 dq \psi ^{\dagger i}_c(-q) \psi ^c_i(-p+q) \psi ^{\dagger j}_b(0) \psi ^a_j(0) \state{0} \nonumber \\ & + {{N^2-1} \over N} \psi ^{\dagger i}_b(-p) \psi ^a _i(0) \state{0}, \eer namely, a color singlet which is a mixture of four quarks and two quarks. The basic feature used in this section has been the fact that $k\gg N_c $. This holds, in fact, not only for large number of fundamental representations but obviously also for any large level gauged WZW $SU(N)$ model. \section{summary} In this work we have calculated the mesonic spectra of several $QCD_2\ $ models by employing bosonization, light-front quantization and expansion in large number of colors or flavors. The main results of the work are a) An approximated spectrum of the ``adjoint fermions" model. b) The exact leading order in ${1\over N_f}$ spectra of multi-flavor fundamental representation. As for (a), our approximation is similar to the one used in the fermionic picture \cite{kutasov}, namely, dropping the terms that mix wave-functions with different number of current creation operators. The physical meaning of such approximation is suppressing pair creation and pair annihilation. This approximation is not quite justified, but it gives us hint about the structure of the spectrum. Obviously, the most urgent task in this direction is to look for methods to solve the full wave equation. The second result states that the spectrum of the large $N_f$ fundamental fermions is built out of a single massive particle with $M^2\sim e^2 N_f $. For comparison see ref.\cite{Engel} where an analysis of similar cases is discussed in the Hamiltonian formalism. In fact, it is a universal behavior of any gauged $SU(N)$ WZW model with $k \gg N$. The interesting question that naturally arise is what happens in the intermediate region, where $k \sim N$. This region can be realized for instance, in the case of multi-flavor $QCD_2$ with $N_f \sim N_c$. It is reasonable to expect that the mesonic spectrum in this region will lie between that of a single massive state and that with an exponential density growth. In the absence of exact analytical methods one may have to invoke numerical diagonalization of the bosonized $M^2$ operator in a similar way to that of the fermionic picture\cite{gyan}. For the analysis of the baryonic spectrum in the bosonization approach it is essential to consider the case of massive quarks\cite{DFS}. This maybe also the case for the mesonic spectrum since the mass term couples the colored, flavored and baryon number sectors. In fact, even for the massless case a better strategy is to solve the massive case and then go to the massless limit. The extraction of the mesonic spectrum in the massive case is much more evolved since the mass term can not be written in a simple fashion in terms of the currents. However, one can systematically expand the mass term in powers of ${m_q\over e}$ where $m_q$ is the quark mass. Solving the wave equations in the presence of these massive perturbation deserves a further future study. Recently, the theories of $YM_2$ and $QCD_2\ $ were analyzed as ``perturbed" topological coset models\cite{FHS}. The spectrum in that approach which was deduced using a BRST procedure includes a peculiar massive state which was not detected in other approaches including the present. This discrepancy maybe related to the different approximation used in that approach. Clearing up this point as well as the implementation of that method to the case of adjoint fermions and other possible generalization deserves a further investigation. Other methods inherited from string theory and conformal field theory can be also be applied to the analysis of the models discussed\cite{abdalla}. $QCD_2\ $ models can be generalized to a much richer class of theories which are also gauge invariant and renormalizable\cite{Doug}. The framework of this generalization is the formulation of the $YM_2$ functional integral in terms of an action which is linear in $F$ and includes an additional auxiliary pseudoscalalr field. Ordinary $QCD_2\ $ has a quadratic term in the latter field while taking any arbitrary function $f$ of this auxiliary field spans the space of generalized models. The analysis presented in the present paper can be applied also to those models. The momentum operator $P^-$ rather then being quadratic in $ {1\over{\partial _- }}J $ it will take the general form of $ f({1\over{\partial _- }}J) $. It will be interesting to compare the outcome of the methods used in the current work to those derived in ref.\cite{Doug}. \bigskip \bigskip {\begin{center} \begin{large We thank for M. Engelhardt, Y. Frishman and D. Kutasov for stimulating conversations. We would specially like to thank S. Yankielowicz for numerous discussions in various part of the work. \section{Appendix} A detailed calculation of currents commutators. We would like to calculate $[P^-,J^{j_1}_{j_2}(-p_1)]$, where $P^-={e^2 \over \pi}\int ^\infty _0 {dp\over {p^2}} J^i_j(-p)J^j_i(p)$. Therefore we would calculate $[\int ^\infty _0 {dp\over {p^2}} J^i_j(-p) J^j_i(p),J^{j_1}_{j_2}(-p_1)]$ by using the affine Lie algebra\eqref{Kac}: \ber \lefteqn{[\int ^\infty _0 {dp\over {p^2}} J^i_j(-p)J^j_i(p), J^{j_1}_{j_2}(-p_1)]=} \\ && \\ && \int ^\infty _0 {dp\over {p^2}} J^i_j(-p)[J^j_i(p),J^{j_1}_{j_2}(-p_1)] + \int ^\infty _0 {dp\over {p^2}} [J^i_j(-p),J^{j_1}_{j_2}(-p_1)]J^j_i(p) = \\ && \\ && \int ^\infty _0 {dp\over {p^2}} J^i_j(-p)\{ p{k\over 2} (\delta ^j_{j_2} \delta ^{j_1}_i -{1\over N}\delta ^j_i \delta ^{j_1}_{j_2})\delta (p-p_1) \\ && +\half (J^{j_1}_i(p-p_1)\delta ^j_{j_2}-J^j_{j_2}(p-p_1)\delta ^{j_1}_i) \} \\ && +\half \int ^\infty _0 {dp\over {p^2}} (J^{j_1}_j(-p-p_1)\delta ^i_{j_2} -J^i_{j_2}(-p-p_1)\delta ^{j_1}_j)J^j_i(p) = \\ && \\ && {k\over 2p_1}J^{j_1}_{j_2}(-p_1) \\ && +\half \int ^\infty _{-p_1} {dp \over {{(p+p_1)}^2}} J^i_{j_2}(-p-p_1)J^{j_1}_i(p) -\half \int ^\infty _{-p_1} {dp \over {{(p+p_1)}^2}} J^{j_1}_j(-p-p_1)J^j_{j_2}(p) \\ && +\half \int ^\infty _0 {dp \over {p^2}} J^{j_1}_j(-p-p_1)J^j_{j_2}(p) -\half \int ^\infty _0 {dp \over {p^2}} J^i_{j_2}(-p-p_1)J^{j_1}_i(p) = \\ && \eer The above expression includes annihilation currents as well creation ones. Separating them one from the other we obtain: \ber \lefteqn{={k\over 2p_1}J^{j_1}_{j_2}(-p_1)} \\ && +\half \int ^\infty _0 dp ({1 \over {{(p+p_1)}^2}} - {1\over {p^2}}) J^i_{j_2}(-p-p_1)J^{j_1}_i(p) \\ && -\half \int ^\infty _0 dp ({1 \over {{(p+p_1)}^2}} - {1\over {p^2}}) J^{j_1}_j(-p-p_1)J^j_{j_2}(p) \\ && +\half \int ^0 _{-p_1} {dp \over {{(p+p_1)}^2}} J^i_{j_2}(-p-p_1)J^{j_1}_i(p) \\ && -\half \int ^0 _{-p_1} {dp \over {{(p+p_1)}^2}} J^{j_1}_j(-p-p_1)J^j_{j_2}(p) = \\ && \\ && {k\over 2p_1}J^{j_1}_{j_2}(-p_1) \\ && +\half NJ^{j_1}_{j_2}(-p_1) \int ^0 _{-p_1} {dp \over {{(p+p_1)}^2}} \\ && +\half \int ^0 _{-p_1} dp ({1 \over {{(p+p_1)}^2}} - {1\over {p^2}}) J^{j_1}_j(p)J^j_{j_2}(-p-p_1) \\ && +\half \int ^\infty _0 dp ({1 \over {{(p+p_1)}^2}} - {1\over {p^2}}) \{ J^i_{j_2}(-p-p_1)J^{j_1}_i(p) - J^{j_1}_i(-p-p_1)J^i_{j_2}(p)\} \\ && \eer In the above expression there are four terms: The first one ${k\over 2p}J(-p)$, does not change the number of currents (it has only one current) and it is proportional to the level $k$. It will play a central role in the large $k$ limit. The second term is similar to the first one, but it is proportional to $N$ and it diverges. The divergent part will be compensated by another divergent term (which arises from the fourth term). The third term include two creation currents. Thus the interaction $P^-$, with the help of the algebra created a current. In our discussion we will ignore this part, for the sake of simplicity. The last term includes annihilation currents, and therefore we should evaluate its commutator with other creation current. \ber \lefteqn{[\half \int ^\infty _0 dp ({1\over {{(p+p_1)}^2}} - {1\over {p^2}}) \{ J^i_{j_2}(-p-p_1)J^{j_1}_i(p) - J^{j_1}_i(-p-p_1)J^i_{j_2}(p)\}, J^{j_2}_{j_3}(-p_2)]=} \\ && \\ && \half \int ^\infty _0 dp ({1\over {{(p+p_1)}^2}} - {1\over {p^2}}) \{ J^i_{j_2}(-p-p_1)[J^{j_1}_i(p),J^{j_2}_{j_3}(-p_2)] \\ && +[J^i_{j_2}(-p-p_1),J^{j_2}_{j_3}(-p_2)]J^{j_1}_i(p) -J^{j_1}_i(-p-p_1)[J^i_{j_2}(p),J^{j_2}_{j_3}(-p_2)] \\ && -[J^{j_1}_i(-p-p_1),J^{j_2}_{j_3}(-p_2)]J^i_{j_2}(p)\} = \\ && \\ && \half \int ^\infty _0 dp ({1\over {{(p+p_1)}^2}} - {1\over {p^2}}) \{ J^i_{j_2}(-p-p_1)p{k\over 2}(\delta ^{j_2}_i \delta ^{j_1}_{j_3} - {1\over N} \delta ^{j_1}_i \delta ^{j_2}_{j_3}) \delta (p-p_2) \\ && \half J^i_{j_2}(-p-p_1)J^{j_2}_i(p-p_2)\delta ^{j_1}_{j_3} -\half J^i_{j_2}(-p-p_1)J^{j_1}_{j_3}(p-p_2) \delta ^{j_2}_i \\ && -\half NJ^i_{j_3}(-p-p_1-p_2)J^{j_1}_i(p) \\ && -J^{j_1}_i(-p-p_1)p{k\over 2}(\delta ^i_{j_3} \delta ^{j_2}_{j_2} - {1\over N} \delta ^i_{j_2} \delta ^{j_2}_{j_3}) \delta (p-p_2) \\ && +\half NJ^{j_1}_i(-p-p_1)J^i_{j_3}(p-p_2) \\ && -\half J^{j_2}_i(-p-p_1-p_2)J^i_{j_2}(p)\delta ^{j_1}_{j_3} +\half J^{j_1}_{j_3}(-p-p_1-p_2)J^i_{j_2}(p)\delta ^{j_2}_i \} = \eer This leads to the following expression: \ber \lefteqn {=-\half p_2 N{k\over 2}({1\over {{(p_1+p_2)}^2}} - {1\over {{p_2}^2}}) J^{j_1}_{j_3}(-p_2-p_1)} \\ && +{1\over 2} \int ^\infty _0 dp ({1\over {{(p+p_1)}^2}} - {1\over {p^2}}) \times \\ && \times \{ J^i_j(-p-p_1)J^j_i(p-p_2)-J^j_i(-p-p_1-p_2)J^i_j(p)\} \delta ^{j_1}_{j_3} \\ && +{N\over 2} \int ^\infty _0 dp ({1\over {{(p+p_1)}^2}} - {1\over {p^2}}) \times \\ && \times \{ J^{j_1}_i(-p-p_1)J^i_{j_3}(p-p_2)-J^i_{j_3}(-p-p_1-p_2)J^{j_1}_i(p)\} \eer A few remarks about the last expression: The first term includes only creation currents, therefore there is no need to evaluate its commutators with other currents. The second term contains creation and annihilation currents, but it is suppressed in the large $N$ limit. The last term is an important term. It is not suppressed at large $N$, and it contains the following expression in it: \beq {N\over 2} \int ^{p_2} _0 dp ({1\over {{(p+p_1)}^2}} - {1\over {p^2}}) J^{j_1}_i(-p-p_1)J^i_{j_3}(p-p_2) \eeq Which may rewritten as: \beq {N\over 2} \int ^{p_2} _0 dp ({1\over {{(p_1+p_2-p)}^2}} - {1\over {{(p_2-p)}^2}}) J^{j_1}_i(p-p_1-p_2)J^i_{j_3}(-p) \eeq This expression leads to the generalized 't Hooft equation.
\section{INTRODUCTION} \label{sec:Intro} The Complex Ginzburg-Landau Equation (CGLE) is the amplitude equation describing universal features of the dynamics of extended systems near a Hopf bifurcation \cite{CrossHohenberg,hohenbergsaarloos}. \begin{equation} \partial_{t} A = a A + ( D_r + {\it i} D_i ) \nabla^{2} A - (b_r + {\it i} b_i ) \mid A \mid^{2} A \ . \label{cgle} \end{equation} Examples of this situation include binary fluid convection \cite{kolodner94}, transversally extended lasers \cite{coullet} and chemical turbulence\cite{kuramoto81}. We will considered here only the one-dimensional case, $A=A(x,t)$, with $x \in [0,L] $. Suitable scaling of the complex amplitude $ A $, space, and time shows that for fixed sign of $a$ there are only three independent parameters in (\ref{cgle}) (with $D_r$ and $b_r>0$ that we assume henceforth). They can be chosen to be $L$, $c_1 \equiv D_i/D_r$, and $c_2 \equiv b_i/b_r$. The CGLE for $a>0$ displays a rich variety of complex spatio-temporal dynamical regimes that have been recently classified in a phase diagram in the parameter space $\{c_1,c_2\}$ \cite{chate1,chate2,chate3}. It is commonly stated that such nontrivial dynamical behavior, occurring also in other nonequilibrium systems, originates from the non-potential or non-variational character of the dynamics \cite{nonvariational}. This general statement needs to be qualified because it involves some confusion in the terminology. For example the term ``non-variational" is often used meaning that there is no Lyapunov functional for the dynamics. But Graham and co-workers, in a series of papers \cite{grahamtel90,grahamtel90a,grahamtel91,graham92,graham93}, have shown that a Lyapunov functional does exist for the CGLE, and they have constructed it approximately in a small-gradient approximation. The correct statement for the CGLE is that it is not a gradient flow. This means that there is no real functional of $A$ from which the right hand side of (\ref{cgle}) could be obtained by functional derivation. Part of the confusion associated with the qualification of ``nonvariational" dynamics comes from the idea that the dynamics of systems having non-trivial attractors, such as limit cycles or strange chaotic attractors, can not be deduced from the minimization of a potential which plays the role of the free energy of equilibrium systems. However, such idea does not preclude the existence of a Lyapunov functional for the dynamics. The Lyapunov functional can have local minima which identify the attractors. Once the system has reached an attractor which is not a fixed point, dynamics can proceed on the attractor due to ``nonvariational" contributions to the dynamical flow which do not change the value of the Lyapunov functional. This just means that the dynamical flow is not entirely determined once the Lyapunov functional is known. This situation is very common and well known in the study of dynamical properties within the framework of conventional statistical mechanics: The equilibrium free energy of the system is a Lyapunov functional for the dynamics, but equilibrium critical dynamics \cite{hohenberg78} usually involves contributions, such as mode-mode coupling terms, which are not determined just by the free energy. The fact that the dynamical evolution is not simply given by the minimization of the free energy is also true when studying the nonequilibrium dynamics of a phase transition in which the system evolves between an initial and a final equilibrium state after, for example, a jump in temperature across the critical point \cite{gunton83}. A Lyapunov functional plays the role of a potential which is useful in characterizing global properties of the dynamics, such as attractors, relative or nonlinear stability of these attractors, etc. In fact, finding such potentials is one of the long-sought goals of nonequilibrium physics \cite{graham89,graham95}, the hope being that they should be instrumental in the characterization of nonequilibrium phenomena through phase transitions analogies. The use of powerful and very general methods based on these analogies has been advocated by a number of authors \cite{ciliberto1,ciliberto2,chate1,chate2,chate3}. In this context, it is a little surprising that the finding of a Lyapunov functional for the CGLE \cite{grahamtel91,graham92,graham93} has not received much attention in the literature. A possible reason for this is that the construction of nonequilibrium potentials has been historically associated with the study of stochastic processes, in particular in the search of stationary probability distributions for systems driven by random noise \cite{graham89,graham95,tira3}. We want to make clear that the finding of the Lyapunov functional for the CGLE \cite{grahamtel91,graham92,graham93}, as well as the whole approach and discussion the present paper is completely within a purely deterministic framework and it does not rely on any noise considerations. A second possible reason for the relative little attention paid to the Lyapunov functional for the CGLE is the lack of any numerical check of the uncontrolled approximations made on its derivation. The main purpose of this paper is precisely to report such numerical check of the results of Graham and collaborators, thus delimiting the range of validity of the approximations involved. We also provide a characterization of the time evolution of the Lyapunov functional in different regions of the phase diagram of the CGLE \cite{chate1,chate2,chate3}, which illustrates the use of such potential. Our main findings are that the expressions by Graham and coworkers behave to a good approximation as a proper Lyapunov potential when phase singularities (vanishing of the modulus of $A$) are not present. This includes non-chaotic regimes as well as states of phase turbulence. In this last case some small but systematic discrepancies with the predictions are found. In the presence of phase singularities the potential is ill-defined and then it is not a correct Lyapunov functional. The paper is organized as follows. For pedagogical purposes, we first discuss in Sect. II a classification of dynamical flows in which notions like relaxational or potential flows are considered. The idea of a potential for the CGLE is clearer in this context. In Sect. III we review basic phenomenology of the CGLE and the main analytical results for the Lyapunov functional of the CGLE. Sections IV and V contain our numerical analyses. Section IV is devoted to the Benjamin-Feir stable regime of the CGLE and Sect. V to the Phase Turbulent regime. Our main conclusions are summarized in Sect.VI. \section{A classification of dynamical flows} \label{pot} \noindent In the following we review a classification of dynamical systems that, although rather well established in other contexts \cite{graham89,graham95}, it is often overlooked in general discussions of deterministic spatio-temporal dynamics. Non-potential dynamical systems are often defined as those for which there is no Lyapunov potential. Unfortunately, this definition is also applied to cases in which there is no {\sl known} Lyapunov potential. To be more precise, let us consider dynamical systems of the general form \begin{equation} \partial_{t} A_i = V_i[A] \label{ds} \end{equation} where $A_i$ represents a set of, generally complex, dynamical variables which are spatially dependent fields: $A_i=A_i({\bf x},t)$. $V_i[A]$ is a functional of them. The notation $A_i^*$ represents the complex conjugate of $A_i$ and for simplicity we will keep the index $i$ implicit. Let us now split $V$ into two contributions: \begin{equation} \label{split} V[A]=G[A]+N[A] \ , \end{equation} \noindent where $G$, the {\sl relaxational} part, will have the form \begin{equation} G[A]=-{\Gamma \over 2} {\delta F[A] \over \delta A^*} \ , \label{G} \end{equation} \noindent with $F$ a real and scalar functional of $A$. $\Gamma$ is an arbitrary hermitic and positive-definite operator (possibly depending on $A$). In the particular case of real variables there is no need of taking the complex conjugate, and hermitic operators reduce to symmetric ones. The functional $N[A]$ in (\ref{split}) is the remaining part of $V[A]$. The important point is that, if the splitting (\ref{split}) can be done in such a way that the following orthogonality condition is satisfied (c.c. denotes the complex conjugate expression): \begin{equation} \int d{\bf x} \left( {\delta F[A] \over \delta A({\bf x})} N[A({\bf x})] + {\rm c.c.} \right)=0\ , \label{preHJ} \end{equation} \noindent then the terms in $N$ neither increase nor decrease the value of $F$, which due to the terms in $G$ becomes a decreasing function of time: \begin{equation} {dF[A({\bf x},t)] \over dt} \le 0 \ . \label{decreasing} \end{equation} If $F$ is bounded from below then it is a Lyapunov potential for the dynamics (\ref{ds}). Equation (\ref{HJ}) with $N=V-G$, that is \begin{equation} \label{HJ} \int d{\bf x} \left( {\delta F[A] \over \delta A({\bf x})} \left( V[A({\bf x})] + {\Gamma \over 2} {\delta F[A] \over \delta A^*({\bf x})} \right) + {\rm c.c.} \right)=0\ , \end{equation} can be interpreted as an equation for the Lyapunov potential $F$ associated to a given dynamical system (\ref{ds}). It has a Hamilton-Jacobi structure. When dealing with systems perturbed by random noise, $\Gamma$ is fixed by statistical requirements, but in deterministic contexts such as the present paper, it can be arbitrarily chosen in order to simplify (\ref{HJ}). Solving (\ref{HJ}) is in general a difficult task, but a number of non-trivial examples of the splitting (\ref{split})-(\ref{decreasing}) exist in the literature. Some of these examples correspond to solutions of (\ref{HJ}) found in the search of potentials for dynamical systems \cite{grahamtel91,grahamtel90,grahamtel90a}. Other examples just correspond to a natural splitting of dissipative and non-dissipative contributions in the dynamics of systems with well established equilibrium thermodynamics, as for example models of critical dynamics \cite{hohenberg78} or the equations of nematodynamics in liquid crystals \cite{sagues87}. Once the notation above has been set-up, we can call relaxational systems those for which there is a solution $F$ of (\ref{HJ}) such that $N=0$, that is all the terms in $V$ contribute to decrease $F$. Potential systems can be defined as those for which there is a nontrivial (i.e. a non-constant) solution $F$ to (\ref{HJ}). In relaxational systems there is no long-time dynamics, since there is no time evolution of $A$ once a minimum of $F$ is reached. On the contrary, for potential systems for which $N\neq 0$, the minima of $F$ define the attractors of the dynamical flow, but once one of these attractors is reached, nontrivial sustained dynamics might exist on the attractor. Such dynamics is determined by $N$ and maintains a constant value for the functional $F$. A possible more detailed classification of the dynamical flows is the following: \begin{itemize} \item[1.-] Relaxational gradient flows: Those dynamical systems for which $N=0$ with $\Gamma$ proportional to the identity operator. In this case the time evolution of the system follows the lines of steepest descent of $F$. A well known example is the so called Fisher-Kolmogorov equation, also known as model A of critical dynamics \cite{hohenberg78}, or (real) Ginzburg-Landau equation for a real field $A({\bf x},t)$: \begin{equation} \label{modela} \dot A = \alpha A + \gamma \nabla^{2} A - \beta \mid A \mid ^{2} A\ , \end{equation} \noindent where $\alpha$, $\gamma$ and $\beta$ are real coefficients. This equation is of the form of Eqs. (\ref{ds})-(\ref{G}) with $N=0$, $\Gamma = 1$, and $F = F_{GL}[A]$, the Ginzburg-Landau free energy: \begin{equation} \label{GL} F_{GL}[A] = \int d{\bf x} \left( -{\alpha}\mid A \mid ^{2} + {\gamma } \mid \nabla A \mid ^{2} + {\beta \over 2} \mid A \mid ^{4} \right) \end{equation} \item[2.-] Relaxational non-gradient flows: Still $N=0$ but with $\Gamma$ not proportional to the identity, so that the relaxation to the minimum of $F$ does not follow the lines of steepest descent of $F$. The matrix operator $\Gamma$ might depend on $A$ or involve spatial derivatives. A well known example of this type is the Cahn-Hilliard equation of spinodal decomposition, or model B of critical dynamics for a real variable $A$. \cite{hohenberg78}: \begin{equation} \label{modelb} \dot A = (- {1 \over 2}\nabla^2) \left( -{\delta F_{GL}[A] \over \delta A} \right) \ , \end{equation} The symmetric and positive-definite operator $(- \nabla^2)$ has its origin in a conservation law for $A$. \item[3.-] Non-relaxational potential flows: $N$ does not vanish, but the potential $F$, solution of (\ref{HJ}) exists and is non-trivial. Most models used in equilibrium critical dynamics \cite{hohenberg78} include non-relaxational contributions, and therefore belong to this category. A particularly simple example is \begin{equation} \label{ANLS} \dot A = -(1+i) {\delta F_{GL}[A] \over \delta A^*} \ , \end{equation} \noindent where now $A$ is a complex field. Notice that we can not interpret this equation as being of type 1, because $(1+i)$ is not a hermitic operator, but still $ F_{GL}$ is a Lyapunov functional for the dynamics. Equation (\ref{ANLS}) is a special case of the Complex Ginzburg- Landau Equation (CGLE), in which $V[A]$ is the sum of a relaxational gradient flow and a nonlinear-Schr\"odinger-type term $N[A]=-i {\delta F_{GL}[A] \over \delta A^*} \ $. The general CGLE\cite{hohenbergsaarloos} is of the form (\ref{modela}) but $A$ is complex and $\alpha$, $\gamma$ and $\beta$ are arbitrary complex numbers. For the special case in which $\frac{Re[\gamma]}{Im[ \gamma ]} = \frac{Re[\beta]}{Im[\beta]}$, as for example in (\ref{ANLS}), the Lyapunov functional for the CGLE is known exactly \cite{tel82}. Such choice of parameters has important dynamical consequences\cite{tira2}. Beyond such special cases, the calculations by Graham and coworkers indicate \cite{graham92,graham93} that the CGLE, a paradigm of complex spatio-temporal dynamics, might be classified within this class of non-relaxational potential flows because a solution of (\ref{HJ}) is found. The difficulty is that the explicit form of the potential is, so far, only known as a uncontrolled small-gradient expansion. \item[4.-] Non-potential flows: Those for which the only solutions $F$ of (\ref{HJ}) are the trivial ones (that is $F=$ constant). Hamiltonian systems as for example the nonlinear Schr\"odinger equation are of this type. \end{itemize} \section{A LYAPUNOV FUNCTIONAL FOR THE CGLE} \label{pheno} It is well known that for $a<0$ the one dimensional CGLE (\ref{cgle}) has $A=0$ as a stable solution, whereas for $a>0$ there are Travelling Wave (TW) solutions of the form \begin{equation} A_k= A_{0}e^{i(k x +\omega t)+\varphi_0} \label{tw} \end{equation} \noindent with $A_{0} = \sqrt{(a - D_r k^{2})/b_r}$, $|k|<\sqrt{a/b_r}$, and $\omega = ( b_i a +D_- k^2)/b_r$. We have introduced \begin{equation} D_{-} \equiv D_r b_i - D_i b_r \ . \label{lalinea} \end{equation} $\varphi_0$ is any arbitrary constant phase. The linear stability of the homegeneus solution ( (\ref{tw}) with $k=0$ ) with respect to long wavelength fluctuations divides the parameter space $ \{ c_1 ,c_2 \}$ in two regions: the Benjamin-Feir (BF) stable and the BF unstable zone. This line is given by \cite{bf1,bf2} \begin{equation} D_{+} \equiv D_r b_r +D_i b_i = 0 , \label{bfline} \end{equation} In the BF unstable region ($D_+<0$) there are no stable TW solutions, while in the BF stable region ($D_+>0$) TW's with a wavenumber $k<k_E$ are linearly stable. For $k>k_E$, TW's become unstable through the long wavelength instability known as the Eckhaus instability \cite{eckhaus1,janiaud1}. The Eckhaus wavenumber $k_E$ is given by \begin{equation} k_E^2 = \frac{a b_r D_+}{D_r (3 D_+ b_r + 2 D_- b_i)} \label{keckhaus} \end{equation} Recent numerical work for $a>0$ and $L$ large \cite{chate1,chate2,chate3,egolf195} has identified regions of the parameter space displaying different kinds of regular and spatio-temporal chaotic behavior (obtained at long times from random initial conditions and periodic boundary conditions), leading to a ``phase diagram" for the CGLE. The five different regions, each leading to a different asymptotic phase, are shown in Fig.\ \ref{fig1} as a function of the parameters $c_1$ and $c_2$ ($a>0$, $L$ large). Two of these regions are in the BF stable zone and the other three in the BF unstable one. One of the main distinctions between the diferent asymptotic phases is in the behavior of the modulus of $A$ at long times. In some regions it never vanishes, whereas in others it vanishes from time to time at different points. A more detailed description of the asymptotic behavior in the different regions is as follows: \begin{enumerate} \item Non-Chaotic region. The evolution here ends in one of the Eckhaus-stable TW solutions for almost all the initial conditions. \item Spatio-Temporal Intermittency region. Despite the fact that there exist stable TW, the evolution from random initial conditions is not attracted by them but by a chaotic attractor in which typical configurations of the field $A$ consist of patches of TW interrupted by turbulent bursts. The modulus of $A$ in such bursts typically touches zero quite often. \item Defect Turbulence. This is a strongly disordered phase in which the modulus of $A$ has a finite density of space-time zeros. In addition the space and time correlation functions have a quasi-exponential decay \cite{chate1,chate2}. \item Phase Turbulence. This is a weakly disordered phase in which $|A(x,t)|$ remains away from zero. The temporal correlations decay slower than exponentially \cite{chate1,chate2}. \item Bi-Chaos region. Depending on the particular initial condition, the system ends on attractors similar to the ones in regions 3, 4, or in a new attractor in which the configurations of $A$ consists of patches of phase and defect turbulence. \end{enumerate} An approximate Lyapunov functional for the CGLE was calculated by Graham and collaborators \cite{graham92,graham93,descalzi}. Earlier attempts to find a Lyapunov functional were based on polynomial expansions\cite{tel82,graham75,walgraef82,walgraef83}, while more recent and successful approaches focussed in solving the Hamilton-Jacobi equation (\ref{HJ}) with $\Gamma=1$ in different ways. This was done first by a minimization procedure involving an action integral\cite{grahamtel90,grahamtel90a,grahamtel91}, and more recently by a more direct expansion method \cite{graham92,graham93,descalzi}. This last method provides also expressions in higher dimensions, but we will restrict here to the one-dimensional case. In any case, the solution involves an uncontrolled gradient expansion around space-independent solutions of the CGLE. Such expansion obviously limits the validity of the result to regions in the phase diagram in which there are not strong gradients. Since the expansion was actually performed in polar coordinates, this excludes the regions in which zeros in the modulus of $A$ are typical, since the phase of $A$ becomes singular there. In particular Spatio-temporal intermittency regimes, Bi-chaos and Defect Turbulence are out of the range of validity of Graham's expansion. The meaningfulness of the potential in the other regions of parameter space remains still an open question because of the uncontrolled small gradient approximations used to calculate it, and calls for some numerical check. In their solution of the Hamilton Jacobi equation, Graham and collaborators find different branches of the Lyapunov functional with expressions valid for different values of the parameters. In particular they identify the BF line (\ref{bfline}) as separating two branches of the solution to (\ref{HJ}). The explicit expressions (obtained with $\Gamma=1$) are given in polar coordinates: \begin{equation} A(x,t) = r(x,t)e^{i \varphi(x,t)} \label{polar} \end{equation} In terms of the amplitude $r$, the phase $\varphi$, and their spatial derivates (denoted as $r_x$, $\varphi_x$,$\varphi_{xx}$, etc.) the Lyapunov functional per unit of length $\Phi \equiv F/L$ was found\cite{graham92,graham93}, for $a<0$: \begin{equation} \Phi = \int \biggl\{ b_{r} r^{4} - 2 a r^{2} + 2 \biggl[ D_{r} + \frac{D_{-} b_{i} r^{4}} {3 ( a - b_{r} r^{2} )^{2} } \biggr] r_{x}^{2} - \frac{ 2 D_{-} r^{3}} {3 ( a - b_{r} r^{2} ) } r_{x} \varphi_{x} + 2 D_{r} r^{2} \varphi_{x}^{2} \biggr\} dx \label{anep0} \end{equation} \noindent We note that even in this relatively simple case $a<0$, the result for $\Phi $ is only approximate and its structure reveals a highly non-trivial dynamics. For $a>0$, in the BF stable region ($D_+>0$) the expression for $\Phi$ results: \begin{eqnarray} \Phi & = & \int \biggl\{ b_{r} r^{4} - 2 a r^{2} \nonumber \\ & + &\biggl[ (A_{1} r + B_{1}/r^{2})r^{2}_{x} + (A_{2} r + B_{2}/r) r_{x} \varphi_{x} + 2 (D_{r} r^{2} - D_{-} b_{i} a / \mid b \mid^{2}b_{r} ) \varphi^{2}_{x} \biggr] \nonumber \\ & + & \biggl[ \frac{D_{-} D_{r} b_{i}}{3 b_{r} \mid b \mid^{2}} \varphi^{4}_{x} + \biggl( - \frac{D_{-}^{2} a}{2 b^{4}_{r} r^{2}} - \frac{D_{-}}{b_{r}^{2}} ( D_{-}/b_{r} + 2 D_{i} ) \ln r + C_{1} \biggr) \varphi^{2}_{xx} + \frac{2 D_{-} D_{r}}{3 b^{2}_{r} \mid b\mid^{2}} (b^{2}_{i} - b^{2}_{r} )\frac{\varphi^{3}_{x} r_{x}}{r} \nonumber \\ & + & \frac{2 D_{-} D_{r} b_{i}}{3 b_{r}^{3} \mid b \mid^{2}} (b^{2}_{i} - 2 b^{2}_{r}) \frac{\varphi^{2}_{x} r^{2}_{x}}{r^{2}} - \frac{4 D_{r}b_{i}D_{-}}{3 b^{3}_{r} r} \biggl( 1 + \frac{ \ln(b_{r} r^{2}/a) }{1 - b_{r} r^{2}/a } \biggr) r_{x} \varphi_{x} \varphi_{xx} \biggr] \biggr\} dx \label{anep1} \end{eqnarray} where \begin{eqnarray} A_{1} &=& 2 (D_{r} + b_{i}D_{-}/3b_{r}^{2}) ,\nonumber \\ A_{2} &=& 2 D_{-}/ b_{r} ,\nonumber \\ B_{1 }&=& \frac{2 D_{-} b_{i} a}{3 b_{r}^{3} \mid b \mid^{2}} (2 b^{2}_{r} - b^{2}_{i} ) ,\nonumber \\ B_{2} &=& \frac{2 D_{-} a}{ b_{r}^{2} \mid b \mid^{2}} ( b^{2}_{r} - b^{2}_{i} ) ,\nonumber \\ \end{eqnarray} Clearly, $\Phi$ is ill-defined when $r=0$. By writing-out the Euler-Lagrange equations associated to the minimization of $\Phi$ the TW solutions (\ref{tw}) are identified as local extrema of $\Phi$. Since they occur in families parametrized by the arbitrary phase $\varphi_0$, the minima associated to the TW of a given $k$ are not isolated points but lay on a one-dimensional closed manifold. The non-variational part of the dynamics ($N$ in (\ref{split})) can be explicitly written-down by substracting $G=-{1 \over 2} {\delta F \over \delta A^*}$ with $F=L\Phi$ to the right-hand-side of (\ref{cgle}). It is seen to produce, when evaluated on the manifold of minima of $\Phi$ with a given $k$, constant motion along it. This produces the periodic time dependence in (\ref{tw}) and identify the TW attractors as limit cycles. The value of $k$ for which the corresponding extrema change character from local minima to saddle points is precisely the Eckhaus wavenumber $k_E$. It is remarkable that, although expression (\ref{anep1}) was obtained in a gradient expansion around the homogeneous TW, their minima identify exactly all the TW's of equation (\ref{cgle}), and their frequencies and points of instability are also exactly reproduced. This gives confidence on the validity of Graham's approximations. It should be stressed however that they are not exact and can lead to unphysical consequences. For instance, the value of the potential $\Phi$ evaluated on a TW of wavenumber $k$ ($|k|<\sqrt{a/b_r}$) is \cite{grahamtel91} \begin{equation} \Phi_{k} \equiv \Phi[A_k] = \frac{2 D_{+} a}{\mid b \mid^2} k^{2} \left( 1 - \frac{k^{2}}{6 k_{E}^{2}} \right) + \Phi_{k=0} \label{fisuk} \end{equation} \noindent where $\Phi_{k=0}= - a^{2}/b_r$. For a range of parameter values this expression gives mathematical sense to the intuitive fact that the closer to zero is $k$ the more stable is the associated TW (because its potential is lower). But for some parameter values the minimal potential corresponds to large wavenumbers close to $\pm \sqrt{a/b_r}$. This is counterintuitive and calls for some numerical test. The test will be described below and it will be shown that the wavenumbers close to $\pm \sqrt{a/b_r}$ are out of the range of validity of the small gradient approximations leading to (\ref{anep1}). We already mentioned in the previous section that the Lyapunov functional for the CGLE is exactly known for special values of the parameters \cite{grahamtel91,graham92,tira2}. This happens for $D_{-}\equiv D_r b_i - D_i b_r = 0$, which lies in the BF-stable region as indicated in Fig.\ \ref{fig1} . In this case it is clear that (\ref{cgle}) can be written as \begin{equation} \dot A = - {1 \over 2}{\delta F_{GL}[A] \over \delta A^*} + i b_i \left(-\mid A \mid^2 + \frac{D_r}{b_r}\nabla^2 \right) A \ , \label{special} \end{equation} \noindent where $ F_{GL}[A]$ is (\ref{GL}) for complex $A$ and with $\alpha=2a$, $ \beta = 2 b_r$, and $\gamma = 2 D_r$. It is readily shown that the term proportional to $b_i$ is orthogonal to the gradient part, so that $F_{GL}$ is an exact solution of (\ref{HJ}) for these values of the parameters, and (\ref{special}) is a relaxational non-gradient flow (see classification in section \ref{pot}). It is seen that the approximate expressions (\ref{anep0}) and (\ref{anep1}) greatly simplify when $D_-=0$ leading both to the same expression: \begin{equation} L \Phi = \int \biggl\{- 2 a r^{2} + b_{r} r^{4} + 2 D_{r} r r^{2}_{x} + 2 D_{r} r^{2} \varphi^{2}_{x} \biggr\} dx \label{anepsp} \end{equation} When expressed in terms of $A$ and $A^*$ it reproduces $F_{GL}$ in (\ref{special}). Thus the gradient expansion turns out to be exact on the line $D_-=0$. In the Benjamin-Feir unstable region ($a>0, D_+<0$) the gradient expansion for $\Phi$ becomes\cite{graham93,descalzi}: \begin{eqnarray} \Phi & = & \int \biggl\{ b_{r} r^{4} - 2 a r^{2} +\biggl[ (A_{1} r + \tilde{B_{1}}/r^{2})r^{2}_{x} + (A_{2} r + \tilde{B_{2}}/r) r_{x} \varphi_{x} + 2 D_{r}( r^{2} -\frac{a}{b_{r}}) \varphi^{2}_{x} \biggr] \nonumber \\ & + & \biggl[ \frac{D_{r}^{2} }{ b_{r} } \varphi^{4}_{x} + \biggl( \frac{b_{r}}{2 a^{2} r^{2}} \biggl( \frac{\tilde{B_{2}}^{2}}{4} + \frac{4 D_{r}^{2} a^{2}}{b_{r}^{2}} \biggr) \bigl(r^{2} - \frac{a}{b_{r}} \bigr) - \frac{A_{2}}{2 b_{r}} \bigl( \frac{A_{2}}{4} + D_{i} \bigr) \ln \bigl( \frac{r^{2} b_{r}}{a} \bigr) + \frac{D_{i}^{2}}{b_{r}} \biggr) \varphi^{2}_{xx} \nonumber \\ & - & \frac{4 D_{r} b_{i} D_{-}}{b_{r}^{3} r} \biggl( 1+ \frac{D_{r}\mid b\mid^{2} + 2 b_{r} D_{+}}{b_{i} D_{-} (1 - \frac{b_{r} r^{2}}{a})} \ln \bigl( \frac{r^{2} b_{r}}{a} \bigr) \biggr) \varphi_{x} r_{x}\varphi_{xx} \nonumber \\ & + & \frac{2 D_{r}}{3 b_{r}^{2} r} ( 5 b_{i} D_{r} + D_{i} b_{r} ) \varphi^{3}_{x} r_{x} + \frac{2 b_{i}D_{r}}{3 b_{r}^{3} r^{2}} ( 7 b_{i}D_{r}+ D_{i} b_{r}) \varphi^{2}_{x} r^{2}_{x} \biggr] \biggr\} dx \label{anep3} \end{eqnarray} where, in addition to the previous definitions \begin{eqnarray} \tilde{B_{1 }}&=& \frac{2 D_{-} b_{i} a}{3 b_{r}^{3} \mid b \mid^{2}} (2 b^{2}_{r} - b^{2}_{i} ) ,\nonumber \\ \tilde{B_{2}} &=& - \frac{2 a}{b_{r}^{2} } (D_{r} b_{i} + D_{i} b_{r} ) , \nonumber \\ \end{eqnarray} It was noted before that this expression can be adequate, at most, for the Phase Turbulent regime, since in the other BF unstable regimes $|A|$ vanishes at some points and instants, so that (\ref{anep3}) is ill-defined. The long time dynamics occurs in the attractor defined by the minima of $\Phi$. The Euler-Lagrange equations associated to the minimization of (\ref{anep3}) lead to a relationship between amplitude and phase of $A$ which implies the well known adiabatic following of the amplitude to the phase dynamics commonly used to describe the phase turbulence regime by a nonlinear phase equation. The explicit form of this relationship is \begin{eqnarray} r^2 & = & \frac{a}{br} - \frac{D_r}{b_r} (\nabla \varphi)^2 - \frac{D_i}{b_r} \nabla^2 \varphi + \frac{b_i D_i^2}{2 a b_r^2} \nabla^4 \varphi + 2 \frac{D_r D_i b_i}{a b_r^2} \nabla \varphi \nabla^3 \varphi \nonumber \\ & +& 2 \frac{b_i D_r^2}{a b_r^2} \nabla \varphi \nabla \nabla^2 \varphi + \left[ \frac{D_r D_i b_i}{a b_r^2} - \frac{ \mid D \mid^2 }{a b_r} \right] (\nabla^2 \varphi)^2 \label{radiab} \end{eqnarray} It defines the attractor characterizing the phase turbulent regime. Dynamics in this attractor follows from the nonrelaxational part $N$ in (\ref{split}). When (\ref{radiab}) is imposed in such nonrelaxational part of the dynamics the generalized Kuramoto-Shivashinsky equation containing terms up to fourth order in the gradients \cite{sakaguchi2} is obtained \cite{graham93,descalzi}. We finally note that in the phase turbulent regime the Lyapunov functional $\Phi$ gives the same value \cite{graham93,descalzi} when evaluated for any configuration satisfying (\ref{radiab}), at least within the small gradient approximation. This corresponds to the evolution on a chaotic attractor (associated to the Kuramoto-Sivashinsky dynamics coming from $N$) which is itself embedded in a region of constant $\Phi$ (the potential plateau \cite{graham95}). This plateau consists of the functional minima of $\Phi$ (\ref{radiab}). All the (unstable) TW are also contained in the same plateau, since they satisfy (\ref{radiab}). \section{NUMERICAL STUDIES OF THE LYAPUNOV FUNCTIONAL IN THE BENJAMIN-FEIR STABLE REGIME} \label{results} We numerically investigate the validity of $\Phi[A]$ in (\ref{anep0}), (\ref{anep1}), and (\ref{anep3}) as an approximate Lyapunov functional for the CGLE. When evaluated on solutions $A(x,t)$ of (\ref{cgle}) it should behave as a monotonously decreasing function of time, until $A(x,t)$ reaches the asymptotic attractor. After then, $\Phi$ should maintain in time a constant value characteristic of the particular attractor. All the results reported here were obtained using a pseudo-spectral code with periodic boundary conditions and second-order accuracy in time. Spatial resolution was typically 512 modes, with runs of up to 4096 modes to confirm the results. Time step was typically $\Delta t = .1$ except when differently stated in the figure captions. Since very small effects have been explored, care has been taken of confirming the invariance of the results with decreasing time step and increasing number of modes. System size was always taken as $L=512$, and always $D_r =1$ and $b_i=-1$, so that $c_1=D_i$ and $c_2=-1/b_r$. When a random noise of amplitude $\epsilon$ is said to be used as or added to an initial condition it means that a set of uncorrelated Gaussian numbers of zero mean and variance $\epsilon^2$ was generated, one for each collocation point in the numerical lattice. \subsection{Negative $a$} \label{amenor0} The uniform state $A =0$ is stable for $a<0$. We start our numerical simulation with a plane wave $A=A_0 e^{ikx}$ of arbitrary wavenumber $k=0.295$ and arbitrary amplitude $A_0=1$ (note that the TW's (\ref{tw}) do not exist for $a<0$), and calculate $\Phi$ for the evolving configurations. In order to have relevant nonlinear effects during the relaxation towards $A=0$ we have chosen a small value for the coefficient of the linear term ($a= -0.01$). The remaining parameters were $D_i=1$ and $b_r=1.25$ ($c_1=1$, $c_2=-0.8$). Despite the presence of non-relaxational terms in (\ref{cgle}), $\Phi$ decreases monotonously (see Fig.\ \ref{fig2}) to the final value $\Phi( t = \infty ) = \Phi[A=0] = 0$ confirming its adequacy as a Lyapunov potential. \subsection{Positive $a$. Benjamin-Feir stable regime} \label{bajobf} We take in this section always $a=1$. Non-chaotic (TW) states and Spatio-Temporal Intermittency are the two phases found below the BF line in Fig.\ \ref{fig1}. We first perform several numerical experiments in the non-chaotic region: A first important case is the one on the line $D_{-} = 0$, for which (\ref{anepsp}) is an exact Lyapunov functional $F_{GL}$. We take $D_i=-1$ and $b_r=1$ ($c_{1}=c_{2} = -1$), on the $D_-=0$ line, and compute the evolution of $\Phi=\frac{ F_{GL}}{L}$ along a solution of (\ref{cgle}), taking as initial condition for $A$ a Gaussian noise of amplitude $\epsilon = 0.01$. Despite of the strong phase gradients present specially in the initial stages of the evolution, and of the presence of non-relaxational terms, $\Phi$ decays monotonously in time (Fig.\ \ref{fig3}). The system evolved towards a TW attractor of wavenumber $k = 0.0245$. The value of $\Phi$ in such state is, from Eq. (\ref{fisuk}), $\Phi_{k=0.0245}=-0.998796$. It is important to notice that our numerical solution for $A$ and numerical evaluation of the derivatives in $\Phi$ reproduce this value within a $0.3 \%$ in the last time showed in Fig.\ ref{fig3}, and continues to approach the theoretical value for the asymptotic attractor at longer times\footnote{If a smaller time step is used greater accuracy is obtained. For example, if the time step is reduced to $0.05$ the value of $\Phi$ is reproduced within $10^{-7} \%$. But this takes quite a long computing time.}. We continue testing the Lyapunov functional for $D_i=1$, $b_r=1.25$ ($c_{1}=1$ ,$c_{2} = -0.8$. This is still in the non-chaotic region but, since $D_-\neq0$, $\Phi$ is not expected to be exact, but only a small gradient approximation. We check now the relaxation back to an stable state after a small perturbation. As initial condition we slightly perturb a TW of Eckhaus-stable wavenumber ($k=0.13<k_{E}$) by adding random noise of amplitude $\epsilon =0.09$. $\Phi$ decays monotonously (Fig.\ \ref{fig4}) from its perturbed value to the value $\Phi_{k=0.13}=-0.796632$ as the perturbation is being washed out, as expected for a good Lyapunov functional. A more demanding situation was investigated for $D_i= -1$ and $b_r = 0.5$ (again in the non-chaotic region, $c_{1} = -1$ and $c_{2} = -2$, and $D_-\neq 0$). Two TW of different wavenumbers ($k_1 = 0.4 , k_2 = 0.08$, both Eckhaus-stable) were joined and the resulting state (see inset in Fig.\ \ref{fig5}) was used as initial condition. The TW of smaller wavenumber advances into the other, in agreement with the idea that it is nonlinearly more stable since it gives a smaller value to the potential. As the difference between the two frequencies is large the speed at which one wave advances onto the other is quite large. The interface between the two TW's contains initially a discontinuity in the gradient of the phase which is washed out in a few integration steps. An important observation is that during the whole process the modulus of $A(x,t)$ never vanishes and then the winding number, defined as \begin{equation} \label{winding} \nu \equiv\int_0^L \nabla \varphi dx \end{equation} remains constant ($\nu=20$) (with periodic boundary conditions $\nu$ is constant except at the instants in which the phase becomes singular, that is when $r=0$). After the TW with the smallest wavenumber completely replaced the other, still a phase diffusion process in which the wave adjusts its local wavenumber to the global winding number occurs. The state (limit cycle) finally reached is a TW of $k = 2\pi\nu/L=0.245$. Despite of the complicated and non-relaxational processes occurring $\Phi$ behaves as a good Lyapunov functional monotonously decreasing from the value $\Phi(t=0) = -1.825$ corresponding to the two-wave configuration to the value $\Phi = -1.863$ of the final attractor (Fig.\ ref{fig5}). It would be interesting, as happens in some relaxational models \cite{chan}, finding some relationship between the speed of propagation of the more stable wave onto the less stable one and the difference in $\Phi$ between the two states. The good behavior of $\Phi$ will be obviously lost if the field $A(x,t)$ vanishes somewhere during the evolution. As the next numerical experiment (for $D_i=1$ and $b_r=1.25$, that is $c_{1} = 1$, $c_{2} =-0.8$) we used as initial condition a small ($\epsilon=0.01$) random Gaussian noise. The system was left to evolve towards its asymptotic state (a TW). Fig.\ \ref{fig6} shows that after a transient $\Phi$ monotonously decreases. During the initial transient it widely fluctuates, increasing and decreasing and loosing then its validity as a Lyapunov functional. This incorrect behavior occurs because during the initial stages $A(x,t)$ is small and often vanishes, changing $\nu$. When $A$ (and then $r$) vanishes the phase and (\ref{anep1}) are ill-defined and out of the range of validity of a small gradient approximation. Note the contrast with the case $D_-=0$ in which the potential is exact and well behaved even when $\nu$ is strongly changing. The particular values of the maxima and minima during the transient in which $\nu$ is changing depend on the spatial and temporal discretization, since it is clear from (\ref{anep1}) that $\Phi$ is ill-defined or divergent when $r$ vanishes. Note that this incorrect behavior of $\Phi$ for $D_-\neq 0$ is not a problem for the existence of a Lyapunov functional, but comes rather from the limited validity of the hypothesis used for its approximate construction. Nevertheless, as soon as the strong gradients disappear $\Phi$ relaxes monotonously to the value $\Phi=-0.79997$, corresponding to the final state, a TW of wavenumber $k = -0.0123$. As another test in the non-chaotic region, for $D_i=-1$ and $b_r=0.5$ ($c_{1} = -1$, $c_{2} = -2$) we use as initial condition an Eckhaus-unstable TW ($k=0.54>k_{E}=0.48$) slightly perturbed by noise. The system evolves to an Eckhaus-stable TW ($k = 0.31$) by decreasing its winding number (initially $\nu = 44$ and finally $\nu = 26$). Fig.\ \ref{fig7} shows the evolution of $\Phi$ from its initial value $\Phi(0) = -1.485$ the final one $\Phi = -1.77$. Although there is a monotonously decreasing baseline, sharp peaks are observed corresponding to the vanishing of $r$ associated with the changes in $\nu$. When $\nu$ finally stops changing, so that $A$ is close enough to the final TW, $\Phi$ relaxes monotonously as in Fig.\ \ref{fig4}. It was explained in Sect. \ref{pheno} that there are parameter ranges in which $\Phi$ is smaller near the boundaries for existence of TW, that is near $k =\pm \sqrt{a/b_r}$, than for the homogeneous TW: $k=0$. This happens for example for $D_i=1$, $b_r=1.25$ ($c_{1} = 1$, $c_{2} = -0.8$). The corresponding function $\Phi_k$ is shown in Fig.\ \ref{fig8}. If this prediction is true, and if $\Phi$ is a correct Lyapunov functional, evolution starting with one of these extreme and Eckhaus-unstable TW would not lead to any final TW, since this would increase the value of the Lyapunov functional. This would imply the existence for this value of the parameters of an attractor different from the TW's perhaps related to the Spatio-Temporal Intermittency phenomenon. We use as initial condition at the parameter values of Fig.\ \ref{fig8} an unstable TW of wavenumber $k = 0.64$ ($\Phi \approx -0.81$), slightly perturbed by noise. From Fig.\ \ref{fig8}, the system should evolve to a state with a value of $\Phi$ value even lower than that. What really happens can be seen in Fig.\ \ref{fig9}. The system changes its winding number from the initial value $\nu=52$, a process during which $\Phi$ widely fluctuates and is not a correct Lyapunov functional, and ends-up in a state of $\nu = 5$, with a value of $\Phi$ larger than the initial one. After this the system relaxes to the associated stable TW of $k=2\pi\nu/L=0.061 < k_{E}= 0.23$. As clearly stated by Graham and coworkers, the expressions for the potential are only valid for small gradients. Since $k$ is a phase gradient, results such as Fig.\ \ref{fig8} can only be trusted for $k$ small enough. Finally, we show the behavior of $\Phi$ in the Spatio-Temporal Intermittency regime. Since $\nu$ is constantly changing in this regime it is clear that (\ref{anep1}) will not be a good Lyapunov functional and this simulation is included only for completeness. We take $D_i=0$ and $b_r=0.5$ ($c_1=0$, $c_2=-2$) and choose as initial condition a TW with $k= 0.44>k_{E}= 0.30$ ($\Phi=-1.89814$), with a small amount of noise added. The TW decreases its winding number and the system reaches soon the disordered regime called Spatio-Temporal Intermittency. Fig.\ \ref{fig10} shows that the time evolution of $\Phi$ is plagued with divergences, reflecting the fact that $\nu$ is constantly changing (see inset). It is interesting to observe however that during the initial escape from the unstable TW $\Phi$ shows a decreasing tendency, and that its average value in the chaotic regime, excluding the divergences, seems smaller than the initial one. \section{NUMERICAL STUDIES OF THE LYAPUNOV FUNCTIONAL IN THE PHASE TURBULENCE REGIME} \label{pt} The Phase Turbulence regime is characterized by the absence of phase singularities (thus $\nu$ is constant). This distinguishes it as the only chaotic regime for which $\Phi$ would be well-defined. Graham and co-workers\cite{graham93,descalzi} derived especially for this region an expression proposed as Lyapunov functional in the small gradient approximation (\ref{anep3}). We recall that the calculations in \cite{graham93,descalzi} predict that the phase turbulent attractor lies on a potential plateau, consisting of all the complex functions satisfying (\ref{radiab}), in which all the unstable TW cycles are also embedded. The value of the potential on such plateau can be easily calculated by substituting in (\ref{anep3}) an arbitrary TW, and the result is \begin{equation} \label{plateau} \Phi_{pl}=-{a^2 \over b_r} \ . \end{equation} We note that this value does not depend on $D_i$ nor $D_r$ and then it is independent of $c_1$, the vertical position in the diagram of Fig.\ \ref{fig1}, within the phase turbulence region. In this section we take also $a=1$. We perform different simulations for $D_i=1.75$ and $b_r=1.25$ ($c_1=1.75$, $c_2=-0.8$). In the first one, we start the evolution with the homogeneous oscillation solution (TW of $k=0$). This solution is linearly unstable, but since no perturbation is added, the system does not escape from it. The potential value predicted by (\ref{plateau}) is $\Phi_{pl}=-0.8$. This value is reproduced by the numerical simulation up to the sixth significant figure for all times (Fig.\ \ref{fig11}, solid line). This agreement, and the fact that the unstable TW is maintained, gives confidence in our numerical procedure. In a second simulation, a smooth perturbation (of the form $\mu e^{iqx}$ with $q=0.049$ and $\mu=0.09$) is added to the unstable TW and the result used as initial condition. This choice of perturbation was taken to remain as much as possible within the range of validity of the small gradient hypothesis. After a transient the perturbation grows and the TW is replaced by the phase turbulence state (the winding number remains fixed to $0$). The corresponding evolution of $\Phi$ is shown in Fig.\ \ref{fig11} (long-dashed line). The value of the potential increases from $\Phi_{pl}$ to a higher value, and then irregularly oscillates around it. Both the departure and the fluctuation are very small, of the order of $10^{-4}$ times the value of $\Phi$. Simulations with higher precisions confirm that these small discrepancies from the theoretical predictions are not an artifact of our numerics, but should be attributed to the terms with higher gradients which are not included in (\ref{anep3}). As a conclusion, the prediction that the phase turbulence dynamics, driven by non-relaxational terms, maintains constant $\Phi$ in a value equal to the one for TW is confirmed within a great accuracy. It is interesting however to study how systematic are the small deviations from the theory. To this end we repeat the launching of the TW with a small perturbation for several values of $D_i=c_1$, for the same value of $b_r$ as before. The prediction is that $\Phi$ should be independent of $c_1$. The inset in Fig.\ \ref{fig11} shows that the theoretical value $\Phi_{pl}=-0.8$ is attained near the BF line, and that as $c_1$ is increased away from the BF line there are very small but systematic discrepancies. The values shown for the potential are time averages of its instantaneous values, and the error bars denote the standard deviation of the fluctuations around the average. Again for $c_1=1.75$, $c_2=-0.8$, we perform another simulation (Fig.\ \ref{fig11}, short-dashed line) consisting in starting the system in a random Gaussan noise configuration, of amplitude $0.01$, and letting it to evolve towards the phase turbulence attractor. As in other cases, there is a transient in which $\Phi$ is ill-defined since the winding number is constantly changing. After this $\Phi$ decreases. This decreasing is not monotonous but presents small fluctuations around a decreasing trend. The decreasing finally stops and $\Phi$ remains oscillating around approximately the same value as obtained from the perturbed TW initial condition. The final state has $\nu=-1$, so that in fact the attractor reached is different from the one in the previous runs ($\nu=0$) but the difference is the smallest possible and the difference in value of the associated potentials can not be distinguished within the fluctuations of Fig.\ \ref{fig11}. These observations confirm the idea of a potential which decreases as the system advances towards an attractor, and remains constant there, but at variance with the cases in the non-chaotic region here the decreasing is not perfectly monotonous, and the final value is only approximately constant. Since the small discrepancies with the theory increase far from the BF line, and since it is known that condition (\ref{radiab}) can be obtained from an adiabatic-following of the modulus to the phase that losses accuracy far from the BF line, one is lead to consider the role of adiabatic following on the validity of $\Phi$ as a potential. To this end we evaluated $\Phi$ along trajectories $A(x,t)$ constructed with the phase obtained from solutions of (\ref{cgle}), but with modulus replaced by (\ref{radiab}), so enforcing the adiabatic following of the modulus to the phase. No significant improvement was obtained with respect to the cases in which the adiabatic following was not enforced since that, in fact, adiabatic following was quite welll accomplished by the solution of (\ref{cgle}). Then it is not the fact that the solutions of (\ref{cgle}) do not fulfill (\ref{radiab}) exactly, but the absence of higher gradient terms in both (\ref{radiab}) and (\ref{anep3}) the responsible for the small failures in the behavior of $\Phi$. Finally, it is interesting to show that the Lyapunov potential $\Phi$ can be used as a diagnostic tool for detecting changes in behavior that would be difficult to monitor by observing the complete state of the system. For example the time at which the phase turbulence attractor is reached can be readily identified from the time-behavior of $\Phi$ in Fig.\ \ref{fig11}. More interestingly it can be used to detect the escape from metastable states. For example, Fig.\ \ref{fig12} shows $\Phi$ for evolution from a Gaussian noise initial condition ($\epsilon=0.01$). $D_i=2$ and $b_r=1.25$ ($c_1=2$, $c_2=-0.8$). The system reaches first a long lived state with $\nu=2$ not too different from the usual phase turbulent state of $\nu=2$. After a long time however the system leaves this metastable state and approaches a more ordered state that can be described \cite{montagne3} as phase turbulent fluctuations around quasiperiodic configurations related to those of \cite{janiaud1}. More details about this state will be described elsewhere \cite{montagne3}. What is of interest here is that from Fig.\ \ref{fig12} one can easily identify the changes between the different dynamical regimes. In particular the decrease in the fluctuations of $\Phi$ near $t \approx 1000$ identifies the jump from the first to the second turbulence regimes. \section{CONCLUSIONS AND OUTLOOK} \label{conclusiones} The validity of the expressions for the Lyapunov functional of the CGLE found by Graham and coworkers has been numerically tested. The most important limitation is that they were explicitly constructed in a approximation limited to small gradients of modulus and phase. This precludes its use for evolution on attractors such that zeros of $r$ and thus phase singularities appear (defect turbulence, bi-chaos, spatio-temporal intermittency). The same problem applies to transient states of evolution towards more regular attractors, if phase singularities appear in this transient (for instance decay of an Eckhaus unstable TW, evolution from random states close to $A=0$, etc.). A major step forward would be the calculation of the Lyapunov potential for small gradients of the real and imaginary components of $A$, which would be a well behaved expansion despite the presence of phase singularities. Apart from this, if changes in winding number are avoided, expressions (\ref{anep0}), (\ref{anep1}), and (\ref{anep3}) display the correct properties of a Lyapunov functional: minima on stable attractors, where non-relaxational dynamics maintains it in a constant value, and decreasing value during approach to the attractor. These properties are completely satisfied in the non-chaotic region of parameter space, even in complex situations such as TW competition, as long as large gradients do not appear. It is remarkable that, although the potential is constructed trough an expansion around the $k=0$ TW, its minima identify exactly the remaining TW, its stability, and the non-relaxational terms calculated by substracting the potential terms to (\ref{cgle}) give exactly their frequencies. In the phase turbulence regime, however, there are small discrepancies with respect to the theoretical predictions: lack of monotonicity in the approach to the attractor, small fluctuations around the asymptotic value, and small discrepancy between the values of the potential of TW's and of turbulent configurations, that were predicted to be equal. All these deviations are very small but systematic, and grow as we go deeper in the phase turbulence regime. They can be fixed in principle by calculating more terms in the gradient expansion. In addition in order to clarify the conceptual status of non-relaxational and non-potential dynamical systems one can ask about the utility of having approximate expressions for the Lyapunov functional of the CGLE. Several applications have been already developped for the case in which (\ref{cgle}) is perturbed with random noise. In particular the stationary probability distribution is directly related to $\Phi$, and in addition barriers and escape times from metastable TW have been calculated \cite{grahamtel91,descalzi}. In the absence of random noise, $\Phi$ should be still useful in stating the nonlinear stability of the different attractors. In practice however there will be limitations in the validity of the predictions, since $\Phi$ has been constructed in an expansion which is safe only near one particular attractor (the homogeneous TW). Once known $\Phi$, powerful statistical mechanics techniques (mean field, renormalization group, etc. ) can in principle be applied to it to obtain information on the static properties of the CGLE (the dynamical properties, as time-correlation functions, would depend also on the non-relaxational terms $N$, as in critical dynamics \cite{hohenberg78}). Zero-temperature Monte Carlo methods can also be applied to sample the phase turbulent attractors, as an alternative to following the dynamical evolution on it. All those promising developments will have to face first with the complexity of Eqs. (\ref{anep0}), (\ref{anep1}), and (\ref{anep3}). Another use of Lyapunov potentials (the one most used in equilibrium thermodynamics) is the identification of attractors by minimization instead of by solving the dynamical equations. In the case of the TW attractors, solving the Euler-Lagrange equations for the minimization of $\Phi$ is in fact more complex than solving directly the CGLE with a TW ansatz. But the limit cycle character of the attractors, and their specific form, is derived, not guessed as when substituting the TW ansatz. For the case of chaotic attractors (as in the phase turbulence regime) minimization of potentials can provide a step towards the construction of inertial manifolds. In this respect it should be useful considering the relationships between the Lyapunov potential of Graham and coworkers and other objects based on functional norms used also to characterize chaotic attractors \cite{doering1,doering2}. \section{ Acknowledgments} We acknowledge very helpful discussions on the subject of this paper with R. Graham. We also acknowledge helpful inputs of E. Tirapegui and R. Toral on the general ideas of nonequilibrium potentials. RM and EHG acknowledge financial support from DGYCIT (Spain) Project PB92-0046. R.M. also acknowledges partial support from the Programa de Desarrollo de las Ciencias B\'asicas (PEDECIBA, Uruguay), the Consejo Nacional de Investigaciones Cient\'\i ficas Y T\'ecnicas (CONICYT, Uruguay) and the Programa de Cooperaci\'on con Iberoam\'erica (ICI, Spain).
\section{Introduction} \footnotetext[1]{Mathematisches Institut, Universit\"at G\"ottingen, Bunsenstra{\ss}e 3-5, 37073 G\"ottingen, Germany \newline email: <EMAIL>} \thispagestyle{empty} In this paper we will suggest a construction for height functions for line bundles on arithmetic varieties. Following the philosophy of \cite{Bost/Gillet/Soule 93} heights should be objects in arithmetic geometry analogous to degrees in algebraic geometry. So let $K$ be a number field, ${\cal O}_{K}$ its ring of integers and ${\cal X} / {\cal O}_{K}$ an arithmetic variety, i.e. a regular scheme, projective and flat over ${\cal O}_{K}$, whose generic fiber $X / K$ we assume to be connected of dimension $d$. Then we have to fix a metrized line bundle $( {\cal T}, \| . \| )$ or, equivalently, its first Chern class $$\stackrel{\wedge}{{c}_{1}} ({\cal T} , \| . \| ) = (T,g_{T}) \in ~ \stackrel{\wedge}{{ \rm CH}^{1}} ({\cal X}) ~~.$$ The height of a line bundle ${\cal L}$ on ${\cal X}$ should be the arithmetic degree of the intersection of $\stackrel{\wedge}{{c}_{1}}({\cal L} )$ with $(T,g_{T})^{d}$. For this a natural hermitian metric has to be chosen on ${\cal L}$. We fix a K\"ahler metric $\omega_{0}$ on ${\cal X} ({\Bbb{C}})$, invariant under complex conjugation $F_{\infty}$, as in \cite{Arakelov 74}. Then it is well known that the condition on the Chern form to be harmonic defines $\| .\|$ up to a locally constant factor. In order to determine this factor we require $$\stackrel{\wedge}{\deg} \Big( \det R \pi_{*} {\cal L}, \| . \|_{Q} \Big) = 0 .$$ Here $\pi : {\cal X} \longrightarrow {\rm Spec}~ {\cal O}_{K}$ is the structural morphism and $\| . \|_{Q}$ is Quillen's metric (\cite{Quillen 85}, \cite{Bismut/Gillet/Soule 88}) at the infinite places of $K$. \begin{thm} {\bf Fact.} {\rm a)} If the Euler characteristic $\chi ({\cal L})$ does not vanish, such a metric exists. \newline {\rm b)} $\stackrel{\wedge}{{c}_{1}} ({\cal L}, \| .\|)$ is uniquely determined up to a summand $(0,C)$, where $C = (C_{\sigma})_{\sigma : K \hookrightarrow \Bbb{C}}$ is a system of constants on $X \times_{{\rm Spec K}, \sigma} {\rm Spec} ~ \Bbb{C}$ with $$\sum_{\sigma : K \hookrightarrow \Bbb{C}} C_{\sigma} = 0 ~~~~~(and {}~~C_{\sigma} = C_{\bar{\sigma}}).$$ \end{thm} \begin{thm} {\bf Fact.} Such $(0,C) \in ~ \stackrel{\wedge}{{\rm CH}^{1}} ({\cal X})$ are numerically trivial. \end{thm} \begin{thm} {\rm Now we can state our fundamental} \newline {\bf Definition.} The {\rm height} of the line bundle ${\cal L}$ is given by $$h_{{\cal T}, \omega_{0}} ({\cal L}) := ~ \stackrel{\wedge}{\deg} ~\pi_{*} \, \Big[ \stackrel{\wedge}{{c}_{1}} ({\cal L}, \| . \|) \cdot (T, g_{T})^{d} \Big] ,$$ where $\| . \|$ is one of the distinguished metrics specified above. \end{thm} \begin{thm}\label{main} {\rm In this paper we will analyze this definition in the case of arithmetic surfaces. Our main result is} \newline {\bf Theorem.} Let ${\cal C} / {\cal O}_{K}$ be a regular projective variety of dimension $2$, flat over ${\cal O}_{K}$ and generically connected of genus $g$, $x \in ({\cal C} \times_{{\rm Spec} {\cal O}_{K}} {\rm Spec} ~ K) (K)$ be a $K$-valued point and $\Theta$ be the Theta divisor on the Jacobian $J = {\rm Pic}^{g} (C)$ (defined using $x$). On $$\coprod_{\sigma: K \hookrightarrow \Bbb{C}} \Big( {\cal C} \times_{{\rm Spec} {\cal O}_{K}, \sigma} {\rm Spec} ~ \Bbb{C} \Big) (\Bbb{C})$$ let $\omega$ be a K\"ahler form invariant under $F_{\infty}$ and normalized by $$\int_{({\cal C} \times_{{\rm Spec} {\cal O}_{K}, \sigma} {\rm Spec} \Bbb{C})(\Bbb{C})} \omega = 1$$ for every $\sigma$. Then, for line bundles ${\cal L} / {\cal C}$, fiber-by-fiber of degree $g$ and of degree of absolute value less than $H$ on every irreducible component of the special fibers of ${\cal C}$ (with some constant $H \in \Bbb{N}$) $$h_{x,\omega} ({\cal L}) = h_{\Theta} ({\cal L}_{K}) + {\rm O}(1) ,$$ where $h_{\Theta}$ is the height on $J$ defined using the ample divisor $\Theta$. \end{thm} \begin{thm} {\bf Remark.} {\rm Another connection between heights on the Jacobian of a curve and arithmetic intersection theory was obtained by Faltings \cite{Faltings 84} and Hriljac \cite{Hriljac 85}. Recently it has been generalized to higher dimensions and higher codimension Chow groups by K\"unnemann \cite{Kunnemann 95}. They can write down an explicit formula for the N\'{e}ron-Tate height pairing on the Jacobian (higher Picard variety) in terms of arithmetic intersection theory. The main point is that they consider line bundles (cycles) algebraically equivalent to zero. So there is no need for them to scale a metric (to specify the infinite part of the arithmetic cycles occuring). Our approach, to the contrary, seems to work best for sufficiently ample algebraic equivalence classes of line bundles. A formal relationship between our approach and the other one is not known to the author.} \end{thm} \begin{thm} {\rm In order to prove the two facts above we will use the following simple} \newline {\bf Lemma.} Let $f: X \longrightarrow Y$ be a smooth proper map of complex manifolds, where $X$ has a K\"ahler structure $\omega$ and $Y$ is connected, and $E$ be a holomorphic vector bundle on $X$. For a hermitian metric $\| . \|$ on $E$ and a constant factor $D > 0$ we have $$ h_{Q,(E,D \cdot \| . \|)} = h_{Q,(E,\| . \|)} \cdot D^{\chi (E)} .$$ \end{thm} {\bf Proof.} The homomorphism \begin{eqnarray*} (E,\| . \|) & \longrightarrow & (E, D \cdot \| . \|) \\ s & \mapsto & \frac{1}{D} \cdot s \end{eqnarray*} is an isometry inducing the isometry \pagebreak \begin{eqnarray*} \Big( \det R\pi_{*} E, h_{Q,(E,\| . \|)} \Big) & \longrightarrow & \Big( \det R\pi_{*} E, h_{Q,(E,D \cdot \| . \|)} \Big) \\ x & \mapsto & D^{-\chi (E)} \cdot x ~~~~~~~~~~~~~~~~~. \end{eqnarray*} \begin{center} $\Box$ \end{center} \begin{thm} {\bf Proof of Fact 1. Existence:} {\rm Multiplication of $\| . \|$ by $D$ will change the Quillen metric by the factor $D^{\chi (E)}$ and therefore $\stackrel{\wedge}{\deg} \det R \pi_{*} E$ by the summand $[K:{\Bbb{Q}}] \chi(E) \log D$. \newline {\bf Uniqueness:} The harmonicity condition and invariance under $F_{\infty}$ determine $\| . \|$ up to constant factors $D_{\sigma} > 0$ for each $\sigma: K \hookrightarrow \Bbb{C}$ with $D_{\sigma} = D_{\bar{\sigma}}$. The scaling condition requires $$\prod_{\sigma: K \hookrightarrow \Bbb{C}} D_{\sigma}^{\chi(E)} = 1$$ or $\sum_{\sigma: K \hookrightarrow \Bbb{C}} \log D_{\sigma} = 0.$ } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\bf Proof of Fact 2.} {\rm Let $(Z,g_{Z}) \in ~ \stackrel{\wedge}{\rm CH}_{1}({\cal X})$. Then \begin{eqnarray*} (0,C) \cdot (Z,g_{Z}) & = & (0,g_{Z} \cdot \omega_{(0,C)} + C \cdot \delta_{Z}) \\ & = & (0,C \cdot \delta_{Z}) ~~~~~~~~. \end{eqnarray*} $Z$ is a zero-cycle on $X$, so $\delta_{Z}$ will have, independently on $\sigma$, always the integral $\deg Z$. Therefore \begin{eqnarray*} \stackrel{\wedge}{\deg} ~ \pi_{*} \Big[ (0,C) \cdot (Z,g_{Z}) \Big] & = & \frac{\scriptstyle 1}{\scriptstyle 2} ~ \sum_{\sigma} ~~ \left[ C_{\sigma} \int_{X \times_{{\rm Spec K}, \sigma} {\rm Spec} ~ {\Bbb{C}} (\Bbb{C})} \delta_{Z} \right] \\ & = & \frac{\scriptstyle 1}{\scriptstyle 2} ~ (\sum_{\sigma} C_{\sigma}) \cdot \deg Z \\ & = & 0 ~~. \end{eqnarray*} } \begin{center} $\Box$ \end{center} \end{thm} \section{Divisors versus points of the Jacobian} \begin{thm} {\rm The remainder of this paper is devoted to the proof of Theorem \ref{main}. So let $C/K$ be a regular proper algebraic curve of genus $g$ with $C(K) \neq \emptyset$. We consider a regular projective model ${\cal C} / {\cal O}_{K}$. Denote by $J = {\rm Pic}^{g}_{C/K}$ the Jacobian of $C$. When $x \in C(K)$ is chosen we have a canonical isomorphism ${\rm Pic}^{g-1}_{C/K} \longrightarrow {\rm Pic}^{g}_{C/K} = J$ and thus the divisor $\Theta$ on $J$. $\Theta$ induces a closed embedding $i^{'}: J \hookrightarrow {\bf P}^{N}_{K}$ and a "naive" height for $K$-valued points of $J$: $$h_{\Theta} (D) := \log ~ \left( \prod_{\nu \in M_{K}} \max \left\{ \| i(D)_{0} \|_{\nu}, ~ \ldots ~ , \| i(D)_{N} \|_{\nu} \right\} \right) ~~.$$ Accordingly $j^{*} (\Theta)$ induces a morphism $i: C^{g} \stackrel{j}{\longrightarrow} J \stackrel{i^{'}}{\longrightarrow} {\bf P}^{N}_{K}$ and a height function $h_{j^{*} (\Theta)}$ for $K$-valued points of $C^{g}$. Here $j$ denotes the natural map sending a divisor to its associated line bundle. A general construction for heights defined by a divisor, the "height machine", is given in [CS, Chapter VI, Theorem 3.3]. The underlying height $h$ for $K$-valued points of ${\bf P}^{N}_{K}$ is a height in the sense of Arakelov theory \cite{Bost/Gillet/Soule 93} as follows: We choose the regular projective model ${\bf P}^{N}_{{\cal O}_{K}} \supseteq {\bf P}^{N}_{K}$. Every \linebreak $K$-valued point $y$ of ${\bf P}^{N}_{K}$ can be extended uniquely to an ${\cal O}_{K}$-valued point $\underline{y}$ of ${\bf P}^{N}_{{\cal O}_{K}}$. Let $\overline{{\cal O} (1)}$ be the hermitian line bundle on ${\bf P}^{N}_{{\cal O}_{K}}$, where the hermitian metrics at the infinite places are given by $$\left\| x_{0} \right\| := \left( 1 + \left| \frac{x_{1}}{x_{0}} \right|^{2} + \ldots + \left| \frac{x_{N}}{x_{0}} \right|^{2} \right)^{- \frac{1}{2}} {}~~~~~~~~~ ({\rm i.e.~~} \left\| x_{i} \right\| := \left( \left| \frac{x_{0}}{x_{i}} \right|^{2} + \ldots + 1 + \ldots + \left| \frac{x_{N}}{x_{i}} \right|^{2} \right)^{- \frac{1}{2}} ) ~~.$$ Then $h = h_{\overline{{\cal O} (1)}}$ is the height defined by $\overline{{\cal O} (1)}$ in the sense of [BoGS, Definition 3.1.; \linebreak formula (3.1.6)]. } \end{thm} \begin{thm} {\bf Remark.} {\rm We need a better understanding of ${\cal O} (j^{*} (\Theta))$. By Riemann's Theorem [GH, Chapter 2, \S 7] one has $\Theta = \frac{1}{(g-1)!} j_{*} ((x) \times C^{g-1})$, where $j: C^{g} \stackrel{p}{\longrightarrow} C^{(g)} \stackrel{c}{\longrightarrow} J$ factors into a morphism finite flat of degree $g!$ and a birational morphism. So $j^{*} (\Theta)$ is an effective divisor containing the summands $\pi_{k}^{*} (x)$, where $\pi_{k}: C^{g} \longrightarrow C$ denotes $k$-th projection. $${\cal O} \Big( j^{*} \left( \Theta \right) \Big) = \bigotimes_{k=1}^{g} \pi_{k}^{*} \Big( {\cal O} (x) \Big) \otimes {\cal O} \Big( p^{*} (R) \Big)$$ Intuitively, the divisor $R$ on $C^{(g)}$ corresponds to the divisors on $C$ moving in a linear system. This can be made precise, but we will not need that here. } \end{thm} \begin{thm} {\bf Remark.} {\rm It is a difficulty that there are no regular projective models available for $J$ and $C^{g}$, such that arithmetic intersection theory does not work immediately. So we follow [BoGS, Remark after Proposition 3.2.1.] and consider a projective (not necessarily regular) model of $C^{g}$, namely ${\cal C}^{g} := {\cal C} \times_{{\rm Spec} {\cal O}_{K}} \ldots \times_{{\rm Spec} {\cal O}_{K}} {\cal C}$. Hereon let $\overline{\cal T}$ be a line bundle extending $\bigotimes_{k=1}^{g} \pi_{k}^{*} {\cal O} (x)$ equipped with a hermitian metric. One has to define a height $h_{\overline{\cal T}}$ induced by $\overline{\cal T}$. Consider more generally a projective (singular) arithmetic variety ${\cal X} / {\cal O}_{K}$ and a hermitian line bundle $\overline{\cal U}$ on $\cal X$. Then there is a morphism $\iota: {\cal X} \longrightarrow P$ into a projective variety $P$ smooth over ${\rm Spec} ~ {\cal O}_{K}$ and a line bundle ${\cal U_{P}}$ on $P$ such that $\iota^{*} (\cal U_{P}) = {\cal U}$ (see [Fu, Lemma 3.2.], cf. [BoGS, Remark 2.3.1.ii)]). We can even choose $\iota$ in such a way that the hermitian metric on $\overline{\cal{U}}$ is a pullback of one on ${\cal U}_{P}$ (e.g. as a closed embedding). $$\iota^{*} \Big( \overline{{\cal U}_{P}} \Big) = \overline{{\cal U}}$$ Then for an ${\cal O}_{K}$-valued point $\underline{y}$ of ${\cal X}$ one defines \begin{eqnarray*} h_{\overline{\cal U}} \Big( \underline{y} \Big) & := & h_{\overline{\cal U}_{P}} \Big( \iota_{*} (\underline{y}) \Big) \\ & = & \stackrel{\wedge}{\deg} \Big( \stackrel{\wedge}{c_{1}} (\overline{{\cal U}_{P}} ) \Big| \iota_{*} (y) \Big) ~~, \end{eqnarray*} where $( . | . )$ denotes the pairing $\stackrel{\wedge}{{\rm CH}^{1}} (P) \times {\rm Z}_{1} (P) \longrightarrow ~ \stackrel{\wedge}{{\rm CH}^{1}} ({\rm Spec} ~ {\cal O}_{K} )_{\Bbb{Q}}$ from [BoGS, 2.3.]. In [BoGS, Remark after Proposition 3.2.1.] independence of this definition of the $\iota$ chosen is shown. In particular it becomes clear at this point that the pairing $\Big( \stackrel{\wedge}{c_{1}} ( . ) \Big| . \Big)$ can be extended to arbitrary (singular) projective arithmetic varieties over ${\cal O}_{K}$ and satisfies the projection formula $$ \Big( \stackrel{\wedge}{c_{1}} (L) \Big| f_{*} (Z) \Big) = \Big( \stackrel{\wedge}{c_{1}} (f^{*} (L)) \Big| Z \Big) ~~ .$$ } \end{thm} \begin{thm} \label{sing} {\bf Remark.} {\rm If ${\cal X} / {\cal O}_{K}$ is a regular arithmetic variety, one has another pairing \begin{eqnarray*} [.,.]: \stackrel{\wedge}{{\rm CH}^{1}} ({\cal X}) ~ \times \stackrel{~^{\scriptstyle{\wedge}}}{{\rm CH}_{1}} ({\cal X}) & \longrightarrow & \stackrel{\wedge}{{\rm CH}^{1}} ({\rm Spec} ~ {\cal O}_{K})_{\Bbb Q} \\ (z, y) & \mapsto & \pi_{*} [z \cdot y] ~~. \end{eqnarray*} We note that also $\Big[ \stackrel{\wedge}{c_{1}} (.) , . \Big]$ can be extended to arbitrary (singular away from the generic fiber) projective arithmetic varieties. One has to represent $y$ by a cycle $(Y, g_{Y})$ and to put $$\Big[ \stackrel{\wedge}{c_{1}} (\overline{\cal U}) , Y \Big] := \Big( \stackrel{\wedge}{c_{1}} (\overline{\cal U}) \Big| Y \Big) + \Big(0, \Big( \int_{{\cal X} ({\Bbb C})} ~ g_{Y} \omega_{\stackrel{\wedge}{c_{1}} (\overline{\cal U})} \Big)_{\sigma: K \hookrightarrow {\Bbb C}} \Big)$$ obtaining a pairing satisfying the projection formula $$\Big[ \stackrel{\wedge}{c_{1}} (f^{*} (\overline{\cal U})) , y \Big] = \Big[ \stackrel{\wedge}{c_{1}} (\overline{\cal U}) , f_{*} (y) \Big]$$ for $f$ proper and smooth on the generic fiber. In particular, independence of the cycle chosen carries over from the regular case. Indeed, concerning a trivial arithmetic $1$-cycle one is automatically reduced to surfaces and resolution of singularities is known for two-dimensional schemes [CS, Chapter XI by M. Artin]. Let $f$ be one. Note that for cycles with $\omega_{y} = 0$ the push-forward $f_{*}$ makes sense for any proper $f$. If $f$ is a proper birational map inducing an isomorphism on the generic fibers one has $f_{*} f^{*} w = w$ for arithmetic one-cycles and therefore \begin{equation} \label{fun} \Big[ \stackrel{\wedge}{c_{1}} (f^{*} (\overline{\cal U})) , f^{*} (w) \Big] = \Big[ \stackrel{\wedge}{c_{1}} (\overline{\cal U}) , w \Big] ~~. \end{equation} This is useful for the special case of a (singular) projective arithmetic surface. There $[ . , . ]$ can be specialized to a pairing $\Big[ \stackrel{\wedge}{c_{1}} (.) , \stackrel{\wedge}{c_{1}} (.) \Big]$ between hermitian line bundles. This one is symmetric. Indeed, formula (\ref{fun}) tells us, that it is enough to show that after pullback by a birational morphism. But for regular arithmetic surfaces symmetry is clear. } \end{thm} \begin{thm} \label{class} {\bf Lemma.} Let ${\cal X} / {\cal O}_{K}$ be a (singular) projective arithmetic variety and $X/K$ its generic fiber which is assumed to be regular. Further, let $D$ be a divisor on $X$ and $\overline{\cal U}$ be a hermitian line bundle extending ${\cal O} (D)$. Then $h_{D} = h_{\overline{\cal U}} + {\rm O} (1)$ for $K$-valued points of $X.$ \newline {\bf Proof.} {\rm There is a very ample line bundle on $X$ that can be extended to ${\cal X}$. So we may assume $D$ to be basepoint-free (very ample). Then the two height functions arise from the situations $$\begin{array}{cccccccccccccc} & & & & \overline{{\cal O} (1)} & ~~ & & ~~ & & & {}~~~~\overline{\cal U}~~~~ & & \overline{{\cal U}_{P}}~~ & \\ & & & & | & ~~ & {\rm and} & ~~ & & & | & & | & \\ {\rm Spec} ~ {\cal O}_{K} & \stackrel{y}{\hookrightarrow} & {\cal X} & - \stackrel{i}{-} \rightarrow & {\bf P}^{N}_{{\cal O}_{K}} & ~~ & & ~~ & {\rm Spec} ~ {\cal O}_{K} & \stackrel{y}{\hookrightarrow} & {\cal X} & \stackrel{\iota}{\longrightarrow} & P & ~~. \end{array}$$ Here $i$ is the rational map defined by an extension ${\cal U}^{'}$ of ${\cal O} (D)$ over ${\cal X}$. In the generic fiber $i$ is defined everywhere. Note that $iy$ is a morphism by the valuative criterion. Of course, it comes from sections of the line bundle $y^{*} {\cal U}^{\prime}$. Note that ${\cal U}^{'}$ is equipped with a hermitian metric induced by that on ${\cal O} (1)$. $\iota: {\cal X} \longrightarrow P$ is a morphism into a smooth scheme as described above. Thus \begin{eqnarray*} h_{D} (y) & = & \stackrel{\wedge}{\deg} \left( \stackrel{\wedge}{c_{1}} \left( \overline{{\cal O} (1) } \right) \Big| (iy)_{*} ({\rm Spec} ~ {\cal O}_{K}) \right) \\ & = & \stackrel{\wedge}{\deg} \left( \stackrel{\wedge}{c_{1}} \left( (iy)^{*} \overline{{\cal O} (1) } \right) \Big| {\rm Spec} ~ {\cal O}_{K} \right) \\ & = & \stackrel{\wedge}{\deg} \left( \stackrel{\wedge}{c_{1}} \left( (iy)^{*} \overline{{\cal O} (1) } \right) \right) \\ & = & \stackrel{\wedge}{\deg} \left( \stackrel{\wedge}{c_{1}} \left( y^{*} \left( \overline{{\cal U}^{'}} \right) \right) \right) \end{eqnarray*} and, correspondingly, \begin{eqnarray*} h_{\overline{\cal T}} (y) & = & \stackrel{\wedge}{\deg} \left( \stackrel{\wedge}{c_{1}} \left( \overline{{\cal U}_{P}} \right) \Big| (\iota y)_{*} ({\rm Spec} ~ {\cal O}_{K}) \right) \\ & = & \stackrel{\wedge}{\deg} \left( \stackrel{\wedge}{c_{1}} \left( y^{*} \left( \overline{\cal U } \right) \right) \Big| {\rm Spec} ~ {\cal O}_{K} \right) \\ & = & \stackrel{\wedge}{\deg} \left( \stackrel{\wedge}{c_{1}} \left( y^{*} \left( \overline{\cal U } \right) \right) \right) ~~~ . \end{eqnarray*} But $\overline{{\cal U}^{'}}$ and $\overline{\cal U}$ coincide as line bundles on the generic fiber. As bundles their difference is some ${\cal O} (E)$ where $E$ is a divisor contained in the special fibers of ${\cal X}$, while the hermitian metrics differ by a continuous, hence bounded, factor. Therefore, the first arithmetic Chern classes of the pullbacks considered differ only at the infinite and a finite number of finite places by bounded summands. } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\bf Remark.} {\rm a) When one considers $L$-valued points instead of $K$-valued ones, where $L$ is a number field with $[L:K] = d$, the error term becomes ${\rm O}(d)$; i. e. there is a constant $C$ such that $$ \Big| h_{D} (x) - h_{\overline{\cal U}} (x) \Big| < C \cdot d$$ for $L$-valued points $x$ of $X$ and an arbitrary number field $L/K$. The reason for that is simply that the number of the critical places occuring grows as ${\rm O}(d)$. \newline b) The lemma can be applied to ${\cal X} = {\cal C}^{g}$ and $D = \sum_{k=1}^{g} \pi_{k}^{*} (x)$, since $\bigotimes_{k=1}^{g} \pi_{k}^{*} {\cal O} (\overline{x})$ extends ${\cal O} (D)$. } \end{thm} \begin{thm} \label{div} {\rm The height defined by an extension ${\cal U}$ of $\bigotimes_{k=1}^{g} \pi_{k}^{*} ({\cal O} (x))$ is understood by the following } \newline {\bf Proposition.} On ${\cal C} / {\cal O}_{K} $ let $\overline{\cal S}$ be the line bundle ${\cal O} (\overline{x})$, where $\overline{x}$ denotes the closure of $x$ in $\cal C$, equipped with a hermitian metric. For $L$-valued points $P = (P_{1}, \ldots ,P_{g})$ of $C^{g}$ we consider the divisor $\underline{P} := (P_{1}) + \ldots + (P_{g})$ on $C$. Then $$ h_{\overline{\cal S}} (\underline{P}) = h_{\overline{\cal U}} (P) + O(d) {}~~.$$ \end{thm} {\bf Proof.} By [BoGS, Proposition 3.2.2.ii)] we may assume ${\cal U} = {\cal O} \Big( \sum_{k=1}^{g} \pi_{k}^{*} (\overline{x}) \Big)$, where $\overline{x}$ is the closure of $x$ in ${\cal C}$. The extensions of $P$ and $\underline{P}$ over $\cal C$ and ${\cal C}^{g}$ will be denoted by $(\overline{P_{1}}) + \ldots + (\overline{P_{g}})$, respectively $(\overline{P_{1}}, \ldots , \overline{P_{g}})$. Then \begin{eqnarray*} h_{\overline{\cal S}} \Big( (\overline{P_{1}}) + \ldots + (\overline{P_{g}}) \Big) & = & \stackrel{\wedge}{\deg} \Big( \stackrel{\wedge}{c_{1}} (\overline{\cal S}) \Big| (\overline{P_{1}}) + \ldots + (\overline{P_{g}}) \Big) \\ & = & \sum_{k=1}^{g} \stackrel{\wedge}{\deg} \Big( \stackrel{\wedge}{c_{1}} (\overline{\cal S}) \Big| (\overline{P_{k}} ) \Big) \\ & = & \sum_{k=1}^{g} \stackrel{\wedge}{\deg} \Big( \stackrel{\wedge}{c_{1}} \Big( \pi_{k}^{*} (\overline{\cal S}) \Big) \Big| (\overline{P_{1}}, ~ \ldots ~ , \overline{P_{g}} ) \Big) ~~~~~~~~~~~{\rm "projection ~formula"}\\ & = & \stackrel{\wedge}{\deg} \left( \stackrel{\wedge}{c_{1}} \left( \bigotimes_{k=1}^{g} \pi_{k}^{*} \left( \overline{\cal S} \right) \right) \bigg| \left( \overline{P_{1}}, ~ \ldots ~ , \overline{P_{g}} \right) \right) {}~~. \end{eqnarray*} But by construction $\bigotimes_{k=1}^{g} \pi_{k}^{*} (\overline{\cal S})$ is the line bundle ${\cal U}$, equipped with a hermitian metric (and by definition the formula $\Big( \stackrel{\wedge}{c_{1}} (\bigotimes_{k} \overline{{\cal L}_{k}}) \Big| Z \Big) = {\displaystyle\sum}_{k} \Big( \stackrel{\wedge}{c_{1}} (\overline{{\cal L}_{k}}) \Big| Z \Big) $ holds in singular case, too). So we have $$h_{\overline{\cal S}} \Big( (\overline{P_{1}}) + \ldots + (\overline{P_{g}}) \Big) = h_{\overline{{\cal U}}^{'}} \Big( (\overline{P_{1}} , ~ \ldots ~ , \overline{P_{g}}) \Big) ~~~,$$ \newpage {}~ \newline where $\overline{{\cal U}}^{'}$ differs from $\overline{\cal U}$ only by the hermitian metric. The claim follows from [BoGS, Proposition 3.2.2.i)]. \begin{center} $\Box$ \end{center} \begin{thm} \label{two} {\bf Corollary.} Let $P \in C^{g}(L)$ and $\underline{P}$ be the associated divisor on $C$. Then $$h_{\Theta} ({\cal O} (\underline{P})) = h_{\overline{\cal S}} (\underline{P}) + h_{R} (p_{*} P) + {\rm O}(d) ~~,$$ where $h_{R}$ denotes the height for points of $C^{(g)}$ defined by $R$. \end{thm} \section{An observation concerning the tautological line \newline bundle} In this section we start analyzing the fundamental definition 1.3. First we will consider only varieties over number fields and forget about integral models. \begin{thm} {\bf Definition.} Let $\Delta$ be the diagonal in $C \times C$. Then $$\underline{\underline{\cal E}} := \bigotimes_{k=1}^{g} \pi_{k,g+1}^{*} \Big( {\cal O} (\Delta) \Big) $$ will be called the {\rm tautological line bundle} on $C^{g} \times C$. Note that the restriction of $\underline{\underline{\cal E}}$ to $\{ (P_{1}, \ldots , P_{g}) \} \times C$ equals ${\cal O} (P_{1} + \ldots + P_{g})$. By construction $\underline{\underline{\cal E}}$ is the pullback of some line bundle $\cal E$, said to be {\rm the tautological} one on $C^{(g)} \times C$. $$\underline{\underline{\cal E}} = (p \times id)^{*} ({\cal E})$$ \end{thm} \begin{thm} {\bf Proposition.} We have $\det R \pi_{*} {\cal E} = {\cal O}_{C^{(g)}} (-R)$. \end{thm} \begin{thm} {\rm This will be a direct consequence of the following} \newline {\bf Lemma.} Let $\underline{\cal E} := {\cal E} \otimes \pi_{C}^{*} ({\cal O} (-x))$ be a tautological line bundle fiber-by-fiber of degree $g-1$. Then $$\det R \pi_{*} \underline{\cal E} = {\cal O}_{C^{(g)}} \Big( -c^{*} (\Theta) \Big) ~~.$$ \end{thm} {\bf Proof.} The canonical map $c: C^{(g)} \longrightarrow J$ is given by ${\cal E}$ using Picard functoriality. So for a tautological line bundle ${\cal M}$, fiber-by-fiber of degree $g$ on $J \times C$, one has $${\cal E} = (c \times id)^{*} {\cal M} \otimes \pi^{*} {\cal H} ~~,$$ where $\cal H$ is a line bundle on $C^{(g)}$. Putting ${\cal M}_{0} := {\cal M} \otimes \pi_{C}^{*} {\cal O} (-x)$, where $\pi_{C}: J \times C \longrightarrow C$ denotes here the canonical projection from $J \times C$, we get $$\underline{\cal E} = (c \times id)^{*} {\cal M}_{0} \otimes \pi^{*} {\cal H} {}~~.$$ It follows \begin{eqnarray*} \det R \pi_{*} \underline{\cal E} & \cong & \det R \pi_{*} \Big[ (c \times id)^{*} {\cal M}_{0} \otimes \pi^{*} {\cal H} \Big] \\ & = & \det R \pi_{*} \Big[ (c \times id)^{*} {\cal M}_{0} \Big] \\ & = & c^{*} \det R \pi_{*} {\cal M}_{0} ~~, \end{eqnarray*} where we first used the projection formula, which is particularly simple here, since line bundles, fiber-by-fiber of degree $g-1$, have relative Euler characteristic $0$, and afterwords noted \pagebreak that the determinant of cohomology commutes with arbitrary base change \cite{Knudsen/Mumford 76}. But by [MB, Proposition 2.4.2] or [Fa, p. 396] we know $\det R \pi_{*} {\cal M}_{0} = {\cal O}_{J} (- \Theta)$. The assertion follows. \begin{center} $\Box$ \end{center} \begin{thm} {\bf Proof of the Proposition.} {\rm The short exact sequence $$0 \longrightarrow \underline{\cal E} \longrightarrow {\cal E} \longrightarrow {\cal E}|_{C^{(g)} \times \{ x \} } \longrightarrow 0 $$ gives \begin{eqnarray*} \det R \pi_{*} {\cal E} & = & \det R \pi_{*} \underline{\cal E} \otimes {\cal O} \left( {\frac{\scriptstyle 1}{\scriptstyle g!}} p_{*} \left( \sum_{k=1}^{g} \pi_{k}^{*} (x) \right) \right) \\ & = & {\cal O} \left( - c^{*} (\Theta) \right) \otimes {\cal O} \left( \frac{\scriptstyle 1}{\scriptstyle g!} p_{*} \left( \sum_{k=1}^{g} \pi_{k}^{*} (x) \right) \right) \\ & = & {\cal O} (-R) ~~. \end{eqnarray*} \begin{center} $\Box$ \end{center} } \end{thm} \begin{thm} {\bf Corollary.} \label{glatt} $\det R \pi_{*} ({\cal E} \otimes \pi^{*} {\cal O} (R)) = {\cal O}_{C^{(g)}}$. \newline {\bf Proof.} {\rm This is the projection formula for the determinant of cohomology.} \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\rm Let $\cal J$ be the N\'eron model of the Jacobian $J$ of $C$. It is smooth over ${\cal O}_{K}$, consequently ${\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}$ is smooth over ${\cal C}$ and therefore regular. We note that any $K$-valued point of $J$ can be extended uniquely to an ${\cal O}_{K}$-valued point of $\cal J$. On $J \times C$ we have a tautological line bundle ${\cal M}$, fiber-by-fiber of degree $g$. ${\cal M}$ can be extended over ${\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}$. For this let ${\cal M} = {\cal O} (D)$ with some Weil divisor $D$ on $J \times C$. Its closure $\overline{D}$ in ${\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}$ is obviously flat over ${\cal O}_{K}$ and therefore it has codimension $1$. We choose the extension ${\cal O} ({\overline{D}})$ and denote it by $\cal M$ again. $\cal M$ is a perfect complex of ${\cal O}_{{\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}}$-modules. For the existence of the Knudsen-Mumford determinant we need that $$\pi: {\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C} \longrightarrow {\cal J}$$ has finite ${\rm Tor}$-dimension. For this there exists a closed embedding ${\cal C} \longrightarrow P$, where $P$ is smooth over ${\cal O}_{K}$. Thus $\pi$ factorizes as $${\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C} \stackrel{i}{\hookrightarrow} {\cal J} \times_{{\rm Spec} {\cal O}_{K}} P \stackrel{{\rm smooth}}{\twoheadrightarrow} {\cal J} ~~.$$ By [SGA 6, Expos\'e III, Proposition 3.6] it is enough to show that $i$ has finite ${\rm Tor}$-dimension. But $i_{*}$ is exact and ${\cal J} \times_{{\rm Spec} {\cal O}_{K}} P$ is regular implying quasi-coherent sheaves have locally finite free resolutions of bounded length. ${\cal M}$ has, relative to $\pi$, Euler characteristic $1$. Therefore $\cal M$ can be changed by an inverse image of a line bundle on ${\cal J}$, trivial on the generic fiber, in such a way, that we are allowed to assume $$\det R \pi_{*} {\cal M} \cong {\cal O}_{\cal J} ~~.$$ } \end{thm} \section{Choosing hermitian metrics continuously depending on moduli space} \setcounter{equation}{1} \begin{thm} {\bf Fact.} On ${\cal M}_{\Bbb{C}}$ there exists a hermitian metric $\underline{h}$ such that for every point $y \in J ({\Bbb{C}})$ the curvature form satisfies $$c_{1} ({\cal M}_{{\Bbb{C}},y}, \underline{h}_{y}) = g \omega$$ on $( \{ y \} \times C) ({\Bbb{C}}) \cong C (\Bbb{C})$. \newline {\bf Proof.} {\rm The statement is local in $C^{g} (\Bbb{C})$ by partition of unity. By the Theorem on cohomology and base change $R^{0} \pi_{*} {\cal M}_{\Bbb{C}} (g-1)$ is locally free and commutes with arbitrary base change. Hence there exists, locally on $J (\Bbb{C})$, a rational section $s$ of $\cal M$ that is neither undefined nor identically zero in any fiber. First we choose an arbitrary hermitian metric $\| . \|$ on ${\cal M}_{\Bbb{C}}$. Then \begin{equation} \omega^{'} := -d_{C} d_{C}^{c} \log \| s \|^{2} \end{equation} defines a smooth $(1,1)$-form on $(J \times C) ({\Bbb{C}}) \backslash {\rm div} (s)$, that is fiber-by-fiber the curvature form to be considered. Since construction (2) is independent of $s$ as soon as it makes sense at a point, we obtain $\omega^{'}$ as a smooth $(1,1)$-form on $(J \times C) (\Bbb{C})$ closed under $d_{C}$ and cohomologous to $g \omega$ on $\{ y \} \times C(\Bbb{C})$ for any $y \in C^{g} (\Bbb{C})$. The setup $\| . \|_{\underline{h}} = f \cdot \| . \|$ gives the equation \begin{equation} \omega^{'} - g \omega = d_{C} d_{C}^{c} \log | f |^{2} ~~. \end{equation} But $d d^{c}$ is an elliptic differential operator on the Riemann surface $C (\Bbb{C})$, so by Hodge theory it permits a Green`s operator $G$ compact with respect to every Sobolew norm $\| . \|_{\alpha}$. Consequently, there exists a solution $f$ of (3) being smooth on $(J \times C) (\Bbb{C})$. } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\rm We note, that $\det R \pi_{*} {\cal M} \cong {\cal O}_{\cal J}$ and the isomorphism is uniquely determined up to units of ${\cal O}_{K}$. Namely, one has ${\rm Aut}_{{\cal O}_{\cal J}} ({\cal O}_{\cal J}) = \Gamma ({\cal J}, {\cal O}_{\cal J}^{*})$ and already $\Gamma (J, {\cal O}_{J}^{*})$ consists of constants only. In particular, there is a unitary section, uniquely determined up to units of ${\cal O}_{K}$, $${\bf 1} \in \Gamma ({\cal J}, \det R \pi_{*} {\cal M}) ~~.$$ } \end{thm} \begin{thm} {\bf Corollary.} Let $R \in \Bbb{R}$. Then, on ${\cal M}_{\Bbb{C}}$ there exists exactly one hermitian metric $h$, such that for every point $y \in J(\Bbb{C})$ the curvature form $c_{1} ({\cal M}_{{\Bbb{C}}, y}, h_{y}) = \omega$ and for the Quillen metric one has $$h_{Q,h} ({\bf 1}) = R ~~,$$ where ${\bf 1} \in \Gamma (\{ y \} , \det R \pi_{*} {\cal M}_{{\Bbb{C}}, y} )$. \newline {\bf Proof.} {\rm Let $\underline{h}$ be the hermitian metric from the preceeding fact. We may replace $\underline{h}$ by $f \cdot \underline{h}$ with $f \in C^{\infty} (J(\Bbb{C}))$ without any effort on the curvature forms, since they are invariant under scalation. As $\cal M$ has relative Euler characteristic $1$, exactly $$h := \frac{R}{h_{Q, \underline{h}} ({\bf 1})} \cdot \underline{h} $$ satisfies the conditions required.} \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\rm We have to consider ${\cal M}_{\Bbb{C}}$ on $J({\Bbb{C}}) \times (\coprod_{\sigma: K \hookrightarrow {\Bbb{C}}} C (\Bbb{C}))$. The metric $h$ on ${\cal M}_{\Bbb{C}}$ has to be invariant under $F_{\infty}$, its curvature form is required to be $g \omega$ and we want to realise \begin{equation} \prod_{\sigma: K \hookrightarrow \Bbb{C}} h_{Q,h} ({\bf 1}) = 1 \end{equation} simultaneously for all $y \in J(\Bbb{C})$. The first is possible since $\omega$ is invariant under $F_{\infty}$ and the corollary above already gives conditions uniquely determining $h$. (4) can be obtained by scalation with a constant factor over all $J({\Bbb{C}}) \times (\coprod_{\sigma: K \hookrightarrow {\Bbb{C}}} C({\Bbb{C}}))$. Altogether, for every line bundle of degree $g$ on $C$ we have found a distinguished hermitian metric and seen that it depends, in some sense, continuously on the moduli space $J$. One obtains } \end{thm} \begin{thm} {\bf Proposition.} \label{Ara} Let $K$ be a number field and $({\cal C} / {\cal O}_{K}, \omega)$ a regular connected Arakelov surface. Then, on the (non proper) Arakelov variety $({\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}, \pi_{C}^{*} \omega)$ there is a hermitian line bundle $\overline{\cal M}$ with the following properties. \newline {\rm a)} $$(c \times id)^{*} ({\cal M} |_{J \times C}) = {\cal E} \otimes \pi^{*} {\cal O} (R)$$ is the modified tautological line bundle found in section 3. \newline {\rm b)} The hermitian metric $h$ on ${\cal M} |_{\coprod_{\sigma: K \hookrightarrow \Bbb{C}} (J \times C) ({\Bbb{C})}}$ is invariant under $F_{\infty}$ and has curvature form $g \omega$. \newline {\rm c)} For any $y \in J(K)$ one has $\overline{\{ y \}} \subseteq {\cal J}$ and $$\stackrel{\wedge}{\deg} \left( \det R \pi_{*} \left( {\cal M} |_{\overline{\{ y \} } \times_{{\rm Spec} {\cal O}_{K}} {\cal C}}, h_{{\cal M},y} \right) , \| . \|_{Q,h} \right) = 0 ~~.$$ {\bf Proof.} {\rm b) and c) are clear. For a) we know ${\cal E} = (c \times id)^{*} ({\cal M}|_{J \times C}) \otimes \pi^{*} {\cal H}$ from 4.3. But $\cal H$ is determined by $\det R \pi_{*} {\cal E} = {\cal O}_{C^{(g)}} (-R)$ and $\det R \pi_{*} ({\cal M}|_{J \times C}) = {\cal O}_{J}$. \begin{center} $\Box$ \end{center} } \end{thm} \section{An integral model of the symmetric power} \begin{thm} {\bf Remark.} {\rm It turns out here to be very inconvenient to work directly with the N\'eron model $\cal J$ of the Jacobian of $C$. When one considers the tautological line bundle ${\cal M} |_{J \times C}$, fiber-by-fiber of degree $g$ on $J \times C$ with $\det R \pi_{*} {\cal M}|_{J \times C} \cong {\cal O}_{J},$ then ${\cal M}|_{J \times C}$ will even have an (up to constant factor) canonical section. $$\pi_{*} {\cal M}|_{J \times C} \cong {\cal O}_{J}$$ But this section is zero over a codimension two subset of $J$ such that one is led to blow up this subset. } \end{thm} \begin{thm} {\bf Lemma.} {\rm a)} $C^{(g)}$ is a projective variety. \newline {\rm b)} The divisor $S := \frac{1}{g!} p_{*} \Big( \sum_{k=1}^{g} \pi_{k}^{*} (x) \Big) = \frac{1}{(g-1)!} p_{*} \Big( (x) \times C^{g-1} \Big)$ "one of the points is $x$" on $C^{(g)}$ is ample. \newline {\bf Proof.} {\rm a) There are at least two good reasons for that. First $C^{(g)}$ is proper as a quotient of the proper variety $C^{g}$ and b) gives an ample line bundle. On the other hand we can give a high-powered argument as follows. $C^{(g)}$ is the Hilbert scheme ${\rm Hilb}_{C/K}^{g}$ by [CS, Chapter VII by J. S. Milne, Theorem 3.13] and this is known to be projective for a long time [FGA, Expos\'e 221, Theorem 3.2]. \pagebreak \newline b) By [EGA III, Proposition 2.6.2] it is enough to show that $p^{*} {\cal O} (S) = \bigotimes_{k=1}^{g} \pi_{k}^{*} {\cal O} (x)$ is an ample line bundle on $C^{g}$, which is obvious. } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\bf Proposition.} The morphism $c: C^{(g)} \longrightarrow J$ is the blow-up of some ideal sheaf ${\cal I} \subseteq {\cal O}_{J}$ with $$c^{-1} {\cal I} = {\cal O} (-NR) ~~,$$ where $N$ is a positive integer. \newline {\bf Proof.} {\rm $c$ is birational and by the lemma it is a projective morphism. So it is a blow-up of some ideal sheaf ${\cal I} \subseteq {\cal O}_{J}$. Going through the lines of the proof of [Ha, Chapter II, Theorem 7.17] one sees that $c_{*} ({\cal O} (NS))$ for $N \gg 0$, up to tensor product with line bundles in order to make them ideal sheaves, can be used as $\cal I$. But $c^{*} {\cal O} (\Theta) = {\cal O} (S) \otimes {\cal O} (R)$ gives \begin{eqnarray*} c_{*} ({\cal O} (NS)) & = & c_{*} \Big( {\cal O} (-NR) \otimes c^{*} {\cal O} (N \Theta) \Big) \\ & = & c_{*} \Big( {\cal O} (-NR) \Big) \otimes {\cal O} (N \Theta) \end{eqnarray*} and therefore ${\cal I} = c_{*} {\cal O} (-NR)$ for some $N \gg 0$. We have a short exact sequence $$0 \longrightarrow {\cal O}_{C^{(g)}} (-NR) \longrightarrow {\cal O}_{C^{(g)}} \longrightarrow {\cal O}_{R_{N}} \longrightarrow 0 ~~,$$ where $R_{N}$ denotes the $N$-th infinitesimal neighbourhood of $R$. It follows exactness of $$0 \longrightarrow c_{*} {\cal O}_{C^{(g)}} (-NR) \longrightarrow {\cal O}_{J} \longrightarrow c_{*} {\cal O}_{R_{N}} ~~.$$ Now the image of ${\cal O}_{J} \longrightarrow c_{*} {\cal O}_{R_{N}}$ is the structure sheaf of the scheme-theoretic image $I_{N}$ of $R_{N}$ in $J$. So ${\cal I} = c_{*} {\cal O}_{C^{(g)}} (-NR) = {\cal I}_{I_{N}} \subseteq {\cal O}_{J}$. But ${\cal O}_{C^{(g)}} / c^{-1} {\cal I}_{I_{N}} = c^{*} ({\cal O}_{C^{g}} / {\cal I}_{I_{N}})$ and therefore $c^{-1} {\cal I} = c^{-1} {\cal I}_{I_{N}}$ is the ideal sheaf of $I_{N} \times_{J} C^{(g)}$ in ${\cal O}_{C^{(g)}}$. $c^{-1} {\cal I}$ is known to be invertible, so $I_{N} \times_{J} C^{(g)}$ is necessarily pure of codimension $1$ and by construction it contains the scheme $R_{N}$. But $R = c^{*} c_{*} S - S$ contains with one point its complete fiber in $c: C^{(g)} \longrightarrow J$. So $I_{N} \times_{J} C^{(g)}$ must be an infinitesimal thickening of $R_{N}$. On the other hand, when one replaces $R$ by $S$ and considers the scheme-theoretic image $\underline{I_{N}}$ of the $N$-th infinitesimal neighbourhood $S_{N}$, then $\underline{I_{N}} \times_{J} C^{(g)} \supseteq I_{N} \times_{J} C^{(g)}$ is a pure codimension $1$ subscheme not containing any thickening of $R_{N}$, but only other additional summands (it corresponds to the divisor $c^{*} (N \Theta)$). So, necessarily $I_{N} \times_{J} C^{(g)} = R_{N}$ and $$c^{-1} {\cal I} = {\cal O} (-NR) ~~.$$ } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\rm Denote by $\tilde{\cal J}$ the normalization of the blow-up of $\cal J$ with respect to some extension $\underline{\cal I}$ of the ideal sheaf $\cal I$ over $\cal J$. } \newline {\bf Facts.} {\rm a)} $\tilde{\cal J}$ is some (singular) arithmetic variety proper over $\cal J$. \newline {\rm b)} It is an integral model of $C^{(g)}$. \newline {\rm c)} On $\tilde{\cal J}$ one has the line bundle $${\cal R} := (c^{-1} \underline{\cal I})^{\vee}$$ extending ${\cal O} (NR)$ for some $N > 0$. \newline {\rm Note that we do not know whether we have an extension of ${\cal O} (R)$ over $\tilde{\cal J}$. } \end{thm} \begin{thm} {\rm On $\tilde{\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}$ we will consider the hermitian line bundle $$\overline{\cal F} := (c \times id)^{*} \overline{\cal M} ~~,$$ where $c: \tilde{\cal J} \longrightarrow {\cal J}$ denotes here the extension of $C^{(g)} \longrightarrow J$ (the blow-down morphism). } \newline {\bf Facts.} {\rm a)} ${\cal F}|_{J \times C} = {\cal E} \otimes {\cal O} (R)$. \newline {\rm b)} One has $\det R \pi_{*} {\cal F} \cong {\cal O}_{\tilde{\cal J}}$. \newline {\rm Note here, $\tilde{\cal J}$ is not regular, so we do not know whether $\pi: \tilde{\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C} \longrightarrow \tilde{\cal J}$ has finite Tor-dimension. Thus $\det R \pi_{*}$ does may be not exist as a functor, but for line bundles, coming by base change from ${\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}$, the definition makes sense. } \end{thm} \begin{thm} {\bf Remark.} {\rm All in all we obtain a decomposition $$\overline{\cal F}^{\otimes N} \cong \overline{\cal K} \otimes \pi^{*} \overline{\cal R}$$ of hermitian line bundles, where ${\cal K}$ extends ${\cal E}^{\otimes N}$, the $N$-th power of the tautological line bundle on $C^{(g)} \times C$. } \end{thm} \begin{thm} {\bf Remark.} {\rm Any line bundle of degree $g$ on $C$ gives an ${\cal O}_{K}$-valued point $y: {\rm Spec} ~ {\cal O}_{K} \longrightarrow {\cal J}$ and ${\rm Spec} ~ {\cal O}_{K} \times_{{\cal J}, y} \tilde{\cal J}$ will be proper over ${\rm Spec} ~ {\cal O}_{K}$. So at least for some finite field extension $L/K$ there will be an ${\cal O}_{L}$-valued point $\underline{y}: {\rm Spec} ~ {\cal O}_{L} \longrightarrow \tilde{\cal J}$ lifting $y$. Proposition \ref{Ara}.c) gives $$\stackrel{\wedge}{\deg} \left( \det R \pi_{*} \left( {\cal F} |_{\overline{\{ y \} } \times_{{\rm Spec} {\cal O}_{K}} {\cal C}}, h_{{\cal F},y} \right) , \| . \|_{Q,h} \right) = 0 ~~.$$ } \end{thm} \section{Decomposition into two summands} \begin{thm} {\rm In this section we will restrict to the case that ${\cal C}$ is {\it semistable}, i. e. $\pi: {\cal C} \longrightarrow {\rm Spec} ~ {\cal O}_{K}$ is smooth up to codimension } $2$. \newline {\bf Lemma.} {\rm a)} $\tilde{\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}$ is a normal scheme. \newline {\rm b)} $\tilde{\cal J}$ is quasi-projective over ${\cal O}_{K}$. \newline {\bf Proof.} {\rm a) $\tilde{\cal J}$ is normal, so $\tilde{\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}^{\rm smooth}$ is normal by [SGA1, Expos\'e I, Corollaire 9.10]. In particular $\tilde{\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}$ is regular in codimension $1$. Further $\pi: \tilde{\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C} \longrightarrow \tilde{\cal J}$ is flat with one dimensional fibers. By [EGA IV, Corollaire 6.4.2] $\tilde{\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}$ is Cohen-Macaulay in codimension $2$. \newline b) $\cal J$ is quasi-projective over ${\cal O}_{K}$ by [CS, Chapter VIII by M. Artin, \S 4] and blow-ups are projective morphisms. } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\bf Remark.} {\rm Let $\underline{y}: {\rm Spec} ~ {\cal O}_{L} \longrightarrow \tilde{\cal J}$ be an ${\cal O}_{L}$-valued point lifting an ${\cal O}_{K}$-valued point \linebreak $y: {\rm Spec} ~ {\cal O}_{K} \longrightarrow {\cal J}$. Then \begin{eqnarray*} h_{x, \omega} \left( {\cal M}|_{y \times_{{\rm Spec} {\cal O}_{K}} {\cal C}} \right) & = & \frac{\scriptstyle 1}{\scriptstyle [L:K] N} \stackrel{\wedge}{\deg} \pi_{*} \left[ \stackrel{\wedge}{c_{1}} \left( \overline{\cal F}^{\otimes N} |_{\underline{y} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}} \right) \cdot (x,g_{x}) \right] \\ & = & \frac{\scriptstyle 1}{\scriptstyle [L:K] N} \stackrel{\wedge}{\deg} \pi_{*} \left[ \stackrel{\wedge}{c_{1}} \left( \overline{\cal K} |_{\underline{y} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}} \right) \cdot (x,g_{x}) \right] \\ & + & \frac{\scriptstyle 1}{\scriptstyle [L:K] N} \stackrel{\wedge}{\deg} \pi_{*} \left[ \stackrel{\wedge}{c_{1}} \left( \pi^{*} \overline{\cal R} |_{\underline{y} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}} \right) \cdot (x,g_{x}) \right] ~~. \end{eqnarray*} We note here, on ${\rm Spec} ~ {\cal O}_{L} \times_{{\rm Spec} ~ {\cal O}_{K}} {\cal C}$ being in general a singular scheme, $\pi_{*}$ of an intersection with an arithmetic Chern class is defined using an embedding into a regular scheme, where the line bundle comes from by base change (Remark \ref{sing}, \cite{Fulton 75}.) The first equality comes from projection formula. Note that $x$ means here an ${\cal O}_{L}$-valued point of ${\rm Spec} ~ {\cal O}_{L} \times_{{\rm Spec} ~ {\cal O}_{K}} {\cal C}$ whose push-forward to $\cal C$ is $[L : K] (x)$. } \end{thm} \begin{thm} {\rm Let us investigate the first summand. We have ${\cal K} |_{C^{(g)} \times C} = {\cal E}^{\otimes N}$ and this line bundle has a canonical section $s$, which can be extended over the finite places. Using this section we obtain the arithmetic cycle $({\rm div} ~ (s), -\log \| s \|^{2})$ representing $$\stackrel{\wedge}{c_{1}} (\overline{\cal K}) \in ~ \stackrel{\wedge}{CH^{1}} \left( \tilde{\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C} \right) ~~.$$ The scheme part ${\rm div} (s)$ of this cycle is an extension of the tautological divisor representing $c_{1} ({\cal E}^{\otimes N}) \in {\rm CH}^{1} (C^{(g)} \times C)$ (whose restriction to $\{ (x_{1}, \ldots, x_{g}) \} \times C$ is $N (x_{1}) + \ldots + N (x_{g})$). So ${\rm div} ~ (s)$ is the closure of that divisor, possibly plus a finite sum of divisors over the finite places. We note, that $\cal K$ is given by that divisor since $\tilde{\cal J} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}$ is normal. Consequently, if $\underline{y}$ restricts to the $L$-valued point corresponding to the divisor $D$ on $C$, then $$c_{1} \left( {\cal K} |_{\underline{y} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}} \right) = \left( \overline{D} \right) + ({\rm correction ~ terms}) ~~,$$ where $\overline{D}$ denotes the closure of $N D$ over $\cal C$ and the correction terms are vertical divisors which (over all the $y$) occur only over a finite amount of finite places. Their intersection numbers with $(x, g_{x})$, i. e. with the line bundle ${\cal O} (x)$, are bounded by ${\rm O} ([L:K])$. The infinite part $f$ of $\stackrel{\wedge}{c_{1}} (\overline{\cal K}) = ({\cal D}, f)$ is a function on $(C^{(g)} \times C) \backslash {\rm div} ~ (s)$ whose pullback to $C^{g} \times C$ satisfies all the assumptions of Lemma \ref{Int}. We obtain \begin{eqnarray*} \frac{\scriptstyle 1}{\scriptstyle [L:K] N} \stackrel{\wedge}{\deg} \pi_{*} \left[ \stackrel{\wedge}{c_{1}} \left( \overline{\cal K} |_{\underline{y} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}} \right) \cdot (x,g_{x}) \right] & = & \frac{\scriptstyle 1}{\scriptstyle [L:K] N} \left[ \stackrel{\wedge}{\deg} \left( \stackrel{\wedge}{c_{1}} \left( \overline{{\cal O} (x)} \right) ~ \Big| {}~ \left( {\cal D} |_{\underline{y} \times_{{\rm Spec} ~ O_{K}} {\cal C}} \right) \right) ~~~ \ldots \right. \\ & ~ & ~~~\ldots ~~~ + \left. \frac{\scriptstyle 1}{\scriptstyle 2} \sum_{\sigma : L \hookrightarrow {\Bbb {C}}} \int_{C ({\Bbb{C}})} f_{D} \omega_{x} \right] \\ & = & \frac{\scriptstyle 1}{\scriptstyle N} h_{\overline{\cal S}} (N D) + {\rm O} (1) \\ & = & h_{\overline{\cal S}} (D) + {\rm O} (1) ~~. \end{eqnarray*} Note, for the first equation we used the symmetry of the intersection form for hermitian line bundles (Remark \ref{sing}). The denominator $[L:K]$ disappears by [BoGS, formula (3.1.8)]. } \end{thm} \begin{thm} {\rm The second summand is simpler. One has \begin{eqnarray*} \stackrel{\wedge}{c_{1}} \left( \pi^{*} \overline{\cal R} |_{\underline{y} \times_{{\rm Spec} {\cal O}_{K}} {\cal C}} \right) \cdot (x, g_{x}) & = & \pi^{*} \stackrel{\wedge}{c_{1}} \left( {\cal R} |_{\underline{y}} \right) \cdot (x, g_{x}) \\ & = & \stackrel{\wedge}{c_{1}} \left( {\cal R} |_{\underline{y}} \right) + \left( 0, g_{x} \omega_{R} (\underline{y}) \right) ~~, \end{eqnarray*} when one identifies $\tilde{\cal J} \times_{{\rm Spec} ~ {\cal O}_{K}} \overline{ \{ x \} }$ with $\tilde{\cal J}$. The integral $\int_{C ({\Bbb C})} ~ g_{x} \omega_{R} ( . )$ depends smoothly on the parameter, in particular it is bounded. So the push-forward of the right summand is bounded by ${\rm O} ([L:K])$. On the other hand $\pi_{*} \stackrel{\wedge}{c_{1}} \left( {\cal R} |_{\underline{y}} \right) = \left( \stackrel{\wedge}{c_{1}} (\overline{\cal R}) ~ \Big| ~ \underline{y} \right) $, where the last term is defined by embedding $\tilde{\cal J}$ into a scheme $P$ smooth and projective over ${\cal O}_{K}$ \cite{Fulton 75}. Note here we use $\tilde{\cal J}$ is quasi-projective. Thus Lemma \ref{class} gives $$\stackrel{\wedge}{\deg} \left( \stackrel{\wedge}{c_{1}} (\overline{\cal R}) ~ \Big| ~ \underline{y} \right) = \stackrel{\wedge}{\deg} \left( \stackrel{\wedge}{c_{1}} (\overline{{\cal R}_{P}}) ~ \Big| ~ \iota_{*} (\underline{y}) \right) = h_{\overline{{\cal R}_{P}}} \left( \iota_{*} (\underline{y}) \right) = h_{R_{P}} \left( \iota_{*} D \right) = h_{R} (D) {}~~,$$ where $D$ is the divisor corresponding to the restriction of $\underline{y}$ to $C^{(g)}$. } \end{thm} \begin{thm} \label{stab} {\rm We obtain} \newline {\bf Proposition.} Assume $\cal C$ is semistable and ${\cal L} = {\cal O} (\overline{D})$, where $\overline{D}$ is the closure of some divisor on $C$. Then Theorem \ref{main} is true. \newline {\bf Proof.} {\rm By Corollary \ref{two} this is now proven for line bundles coming by restriction from $\cal M$. This way one can realize the line bundles ${\cal O} (D)$ on the generic fiber $C$ for arbitrary divisors $D$ (defined over $K$) of degree $g$ over $C$. Consider the degrees $$\deg {\cal M} |_{y \times_{{\rm Spec} {\cal O}_{K}} {\cal C}_{{\goth p}, i}}$$ for ${\cal O}_{K}$-valued points $y$ of $\cal J$, where ${\cal C}_{{\goth p}, i}$ denote the irreducible components of the special fiber ${\cal C}_{\goth p}$. They are even defined for $\overline{{\cal O} / {\goth p}}$-valued points, where the bar denotes algebraic closure here, and are locally constant over the special fiber ${\cal J}_{\goth p}$. In particular they are bounded since the N\'eron model of an abelian variety is of finite type. The Proposition follows from Lemma \ref{degr}. } \begin{center} $\Box$ \end{center} \end{thm} \section{End of the proof} \begin{thm} \label{blow} {\bf Lemma.} {\rm Let ${\cal C} / {\cal O}_{K}$ be a regular projective arithmetic surface and $p: \tilde{\cal C} \longrightarrow {\cal C}$ be a blow-up of one point. Then $$h_{x, \omega} ({\cal L}) = h_{x, \omega} (p^{*} {\cal L}) ~~,$$ where $x \in {\cal C} ({\cal O}_{K}) = \tilde{\cal C} ({\cal O}_{K})$ and $\cal L$ is a line bundle with $\chi ({\cal L}) \neq 0$. \newline {\bf Proof.} Obviously $p_{*} p^{*} {\cal L} = {\cal L}$ and [SGA6, Expos\'e VII, Lemma 3.5] gives $R^{i} p_{*} p^{*} {\cal L} = 0$ for $i \geq 1.$ In particular $R p_{*} (p^{*} {\cal L}) = {\cal L}$, $R (\pi p)_{*} (p^{*} {\cal L}) = R \pi_{*} {\cal L}$ and $\det R (\pi p)_{*} (p^{*} {\cal L}) = \det R \pi_{*} {\cal L}$. \linebreak This means that $\cal L$ and $p^{*} {\cal L}$ get identical distinguished metrics and therefore \linebreak $\stackrel{\wedge}{c_{1}} \left( p^{*} {\cal L}, \| . \|_{p^{*} {\cal L}} \right) = p^{*} \stackrel{\wedge}{c_{1}} \left( {\cal L}, \| . \|_{{\cal L}} \right) $. On the other hand $p^{*} (x, g_{x}) = (x, g_{x}) + ({\rm exceptional ~ divisor})$, but an exceptional divisor intersects trivially with cycles coming from downstairs. Consequently, \begin{eqnarray*} h_{x, \omega} (p^{*} {\cal L}) & = & \stackrel{\wedge}{\deg} (\pi p)_{*} \left[ p^{*} \stackrel{\wedge}{c_{1}} ({\cal L}, \| . \|_{\cal L}) \cdot p^{*} (x, g_{x}) \right] \\ & = & \stackrel{\wedge}{\deg} (\pi p)_{*} p^{*} \left[ \stackrel{\wedge}{c_{1}} ({\cal L}, \| . \|_{\cal L}) \cdot (x, g_{x}) \right] \\ & = & \stackrel{\wedge}{\deg} \pi_{*} \left[ \stackrel{\wedge}{c_{1}} ({\cal L}, \| . \|_{\cal L}) \cdot (x, g_{x}) \right] ~~~~~~~~~~~~~~~~~~~~{\rm "projection ~ formula"}\\ & = & h_{x, \omega} ({\cal L}) ~~. \end{eqnarray*} } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\bf Corollary.} \label{model} {\rm (Change of model.)} \newline Let ${\cal C}_{1}, {\cal C}_{2} / {\cal O}_{K}$ be two regular projective models of the curve $C/K$ of genus $g$. Then, for divisors $D$ of degree $g$ on $C$ $$h_{x, \omega} \left( {\cal O}_{{\cal C}_{1}} (\overline{D}) \right) = h_{x, \omega} \left( {\cal O}_{{\cal C}_{2}} (\overline{D}) \right) + {\rm O} (1) {}~~,$$ where $\overline{D}$ denotes the closure of $D$ in ${\cal C}_{1}$, respectively ${\cal C}_{2}.$ \newline {\bf Proof.} {\rm By [Li, Theorem II.1.15] one is reduced to the case of the blow-up of one point $p: {\cal C}_{2} \longrightarrow {\cal C}_{1}$. By Lemma \ref{blow} we have to bound the difference $h_{x, \omega} \left( {\cal O}_{{\cal C}_{2}} (\overline{D}) \right) - h_{x, \omega} \left( p^{*} {\cal O}_{{\cal C}_{1}} (\overline{D}) \right)$. $\overline{D}$ will meet the point blown up $i$ times ($0 \leq i \leq g$). We get an exact sequence $$0 \longrightarrow {\cal O}_{{\cal C}_{2}} (\overline{D}) \longrightarrow p^{*} {\cal O}_{{\cal C}_{1}} (\overline{D}) \longrightarrow {\cal O}_{E^{i}} \longrightarrow 0 ~~,$$ where $E$ is the exceptional curve and $E^{i}$ denotes its $i$-th infinitesimal neighbourhood. But now the assertion is a direct consequence of Lemma \ref{red}. } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\bf Lemma.} {\rm (Change of base field.)} \newline Let ${\cal C} / {\cal O}_{K}$ be a regular arithmetic surface, generically of genus $g$, $L/K$ a finite field extension and $$p: {\cal C}^{'} = \overline{{\cal C} \times_{{\rm Spec} ~ {\cal O}_{K}} {\rm Spec} ~ {\cal O}_{L}} \longrightarrow {\cal C}$$ be some resolution of singularities of the base change to ${\cal O}_{L}$. Then, for divisors $D$ of degree $g$ on $C = {\cal C} \times_{{\rm Spec} ~ {\cal O}_{K}} {\rm Spec} ~ K$, $$h_{x, \omega} \left( {\cal O}_{{\cal C}^{'}} (\overline{p^{*} D}) \right) = [L:K] \cdot h_{x, \omega} \left( {\cal O}_{\cal C} (\overline{D}) \right) + {\rm O} (1) ~~.$$ \newline {\bf Proof.} {\rm $p$ is a composition of blow-ups and finite morphisms [CS, Chapter XI by M. Artin]. \linebreak Using the first formulas in the proof of Lemma \ref{blow} successively we obtain \linebreak $R p_{*} p^{*} {\cal O} (\overline{D}) = {\cal O} (\overline{D})$ and $\det R (\pi p)_{*} p^{*} \left( {\cal O} (\overline{D}) \right) = \det R \pi_{*} {\cal O} (\overline{D})$ such that ${\cal O} (\overline{D})$ and $p^{*} {\cal O} (\overline{D})$ get identical distinguished metrics. Here it follows $$h_{x, \omega} \left( p^{*} {\cal O} (\overline{D}) \right) = [L:K] \cdot h_{x, \omega} \left( {\cal O} (\overline{D}) \right) ~~,$$ since $p$ is a morphism of degree $[L:K]$ and the projection formula gives $p_{*} p^{*} Z = [L:K] Z$. $p^{*} {\cal O} (\overline{D})$ and ${\cal O} (\overline{p^{*} D})$ differ by a limited combination of the exceptional divisors such that the assertion follows from Lemma \ref{red}. } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} \label{hori} {\bf Proposition.} For line bundles ${\cal L} = {\cal O} (\overline{D})$, where $\overline{D}$ is the closure of some divisor of degree $g$ on $C$, Theorem \ref{main} is true. \newline {\bf Proof.} {\rm By [AW, Corollary 2.10] there is a stable model for $C \times_{{\rm Spec} ~ K} {\rm Spec} ~ L$ after some finite field extension $L/K$. The assertion follows from Proposition \ref{stab}. } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\bf Proposition.} Theorem \ref{main} is true. \newline {\bf Proof.} {\rm This is a direct consequence of Proposition \ref{hori} and Lemma \ref{degr}. } \begin{center} $\Box$ \end{center} \end{thm} \section{Some technical Lemmata} \begin{thm} \label{fibe} {\bf Lemma.} {\rm (Fibers do not change the height.)} \newline If $\cal L$ is a line bundle on ${\cal C} / {\cal O}_{K}$ with $\chi ({\cal L}) \neq 0$, then $$h_{x, \omega } \left( {\cal L} \otimes {\cal O} (\goth{p}) \right) = h_{x, \omega } ({\cal L})$$ for every prime ideal $\goth{p} \subseteq {\cal O}_{K}$. \newline {\bf Proof.} {\rm One has ${\cal O} (\goth{p}) = \pi^{*} (\goth{p}^{-1})$, hence by projection formula $$\det R \pi_{*} \left( {\cal L} \otimes {\cal O} (\goth{p}) \right) \cong \det R \pi_{*} {\cal L} \otimes {\cal O} ({\goth p})^{- \chi ({\cal L})} ~~.$$ Let $\| . \|$ be one of the distinguished metrics on the line bundle ${\cal L}_{\Bbb{C}}$ on $\coprod_{\sigma: K \hookrightarrow \Bbb{C}} C(\Bbb{C})$. \linebreak We put $\| . \|_\goth{p} = C \cdot \| . \|$ for a distinguished hermitian metric on $({\cal L} \otimes {\cal O} (\goth{p}))_{\Bbb{C}} = {\cal L}_{\Bbb{C}}$. It follows \linebreak $h_{Q, \det R \pi_{*} ({\cal L} \otimes {\cal O} (\goth{p}))} = C^{\chi ({\cal L})} \cdot h_{Q, \det R \pi_{*} {\cal L}}$ and \begin{eqnarray*} \stackrel{\wedge}{\deg} \left( \det R \pi_{*} ({\cal L} \otimes {\cal O} (\goth{p})), h_{Q, \det R \pi_{*} ({\cal L} \otimes {\cal O} (\goth{p}))} \right) & = & \stackrel{\wedge}{\deg} \Big( \det R \pi_{*} {\cal L}, h_{Q, \det R \pi_{*} {\cal L}} \Big) \\ & + & \chi ({\cal L}) \Big[ [K : {\Bbb{Q}}] \log C - \log (\sharp {\cal O} / \goth{p}) \Big] ~~. \end{eqnarray*} Thus a distinguished hermitian metric on $({\cal L} \otimes {\cal O} (\goth{p}))_{\Bbb{C}}$ can be given by $\| . \|_\goth{p} = ( \sharp {\cal O} / \goth{p})^{\frac{1}{[K : {\Bbb Q}]}} \cdot \| . \|$ and it follows $$\stackrel{\wedge}{c_{1}} ({\cal L} \otimes {\cal O} (\goth{p}), \| . \|_\goth{p}) = ~ \stackrel{\wedge}{c_{1}} ({\cal L}, \| . \|) + \pi^{*} \left( \goth{p} ; - \frac{\scriptstyle 2}{\scriptstyle [K : \Bbb{Q}]} \log (\sharp {\cal O} / \goth{p}) ,\ldots ,- \frac{\scriptstyle 2}{\scriptstyle [K : \Bbb{Q}]} \log (\sharp {\cal O} / \goth{p}) \right) ~~.$$ But the arithmetic cycle $\left( \goth{p} ; - \frac{2}{[K : \Bbb{Q}]} \log (\sharp {\cal O} / \goth{p}), \ldots, - \frac{2}{[K : \Bbb{Q}]} \log (\sharp {\cal O} / \goth{p}) \right) \in {}~ \stackrel{\wedge}{{\rm CH}^{1}}({\rm Spec} ~ {\cal O}_{K})$ vanishes after multiplication with the class number $\sharp {\rm Pic} ~ ({\rm Spec} ~ {\cal O}_{K})$, hence it is torsion and therefore numerically trivial. } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} \label{red} {\bf Lemma.} Let $F$ be some vertical divisor on ${\cal C} / {\cal O}_{K}$. Then, for line bundles ${\cal L} / {\cal C}$, fiber-by-fiber of degree $g$, $$h_{x, \omega} ({\cal L} (F)) = h_{x, \omega} ({\cal L}) + {\rm O} (1) ~~.$$ {\bf Proof.} {\rm By Lemma \ref{fibe} we may assume that $E := -F$ is effective. Using induction we are reduced to the case $E$ is an irreducible curve. We have a short exact sequence $$0 \longrightarrow {\cal L} (F) \longrightarrow {\cal L} \longrightarrow {\cal L}_{E} \longrightarrow 0$$ inducing the isomorphism $$\det R \pi_{*} {\cal L} (F) \cong \det R \pi_{*} {\cal L} \otimes (\det R \pi_{*} {\cal L}_{E})^{\vee} ~~.$$ But $\det R \pi_{*} {\cal L}_{E}$ depends only on the Euler characteristic of ${\cal L}_{E}$ and for the degree of that bundle there are only $g+1$ possibilities. So up to numerical equivalence there are only $g+1$ possibilities for $$\stackrel{\wedge}{c_{1}} \left( {\cal L} (F), \| . \|_{{\cal L} (F)} \right) - \stackrel{\wedge}{c_{1}} \Big( {\cal L}, \| . \|_{\cal L} \Big) ~~,$$ where $\| . \|_{\cal L}$ and $\| . \|_{{\cal L} (F)}$ denote distinguished hermitian metrics. } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} \label{degr} {\bf Lemma.} Consider line bundles ${\cal L}$, generically of degree $g$ on ${\cal C}$, equipped with a section $s \in \Gamma (C, {\cal L}_{C})$ over the generic fiber, and assume the degrees $\deg {\cal L} |_{{\cal C}_{{\goth p}, i}}$ of the restrictions of $\cal L$ to the irreducible components of the special fibers to be fixed. Then $$h_{x, \omega} ({\cal L}) = h_{x, \omega} \left( {\cal O} \left( \overline{{\rm div} (s)} \right) \right) + {\rm O} (1) ~~.$$ {\bf Proof.} {\rm We have ${\cal L} = {\cal L}^{'} (E)$, where ${\cal L}^{'} = {\cal O} \left( \overline{{\rm div} (s)} \right)$ is a line bundle induced by a horizontal divisor and $E$ is a vertical divisor. By Lemma \ref{fibe} we may assume $E$ to be concentrated in the reducible fibers of $\cal C$. So, using induction, let $E$ be in one such fiber ${\cal C}_{{\goth p}}$. Then for the degrees $\deg {\cal O} (E) |_{{\cal C}_{{\goth p}, i}}$ there are only finitely many possibilities. But by [Fa, Theorem 4.a)] the intersection form on ${\cal C}_{\goth p}$ \pagebreak is negative semi-definite where only multiples of the fiber have square $0$. \linebreak Hence, for $E$ there are only finitely many possibilities up to addition of the whole fiber, which does not change the height. Lemma \ref{red} gives the claim. } \begin{center} $\Box$ \end{center} \end{thm} \begin{thm} {\bf Lemma.} \label{Int} Let $X$ be a compact Riemann surface and $g \in \Bbb{N}$ be a natural number. Denote by $\Delta$ the diagonal in $X \times X$, by $\delta_{M}$ the $\delta$-distribution defined by $M$ and by $\pi_{i}: X^{g} \times X \longrightarrow X$ (resp. $\pi_{i,g+1}: X^{g} \times X \longrightarrow X \times X$) the canonical projection on the $i$-th component (resp. to the product of the $i$-th and $(g+1)$-th component.) Further let $$f: (X^{g} \times X) \backslash \bigcup_{i=1}^{g} \pi_{i,g+1}^{-1} (\Delta) \longrightarrow {\Bbb{C}}$$ be a smooth function such that the restriction of $$-d_{X} d_{X}^{c} f + \delta_{\Delta} \circ \pi_{1,g+1} + \ldots + \delta_{\Delta} \circ \pi_{g,g+1} = \rho ~~,$$ to $\{ (x_{1}, \ldots, x_{g}) \} \times X$ is a smooth $(1,1)$-form smoothly varying with $(x_{1}, \ldots, x_{g})$. Let $\omega$ be a smooth $(1,1)$-form on $X$. Then $$\int_{X} f(x_{1}, \ldots ,x_{g}, \cdot) \omega$$ depends smoothly on $(x_{1}, \ldots, x_{g}) \in X^{g}.$ \newline {\bf Proof.} {\rm Without restriction we may assume $\int_{X} \omega = 1$. Then, for any $x \in X$ there exists a function $h \in C^{\infty} (X \backslash \{ x \})$, having a logarithmic singularity in $x$, such that $\omega = -dd^{c} h + \delta_{x}$. It follows \begin{eqnarray*} \int_{X} f(x_{1}, \ldots, x_{g}, \cdot) \omega & = & -\int_{X} f(x_{1}, \ldots, x_{g}, \cdot) dd^{c} h + f(x_{1}, \ldots, x_{g}, x) \\ & = & -\int_{X} \Big( d_{X} d_{X}^{c} f(x_{1}, \ldots ,x_{g}, \cdot) \Big) h + f(x_{1}, \ldots, x_{g}, x) \\ & = & \int_{X} \rho (x_{1}, \ldots, x_{g}, \cdot) h - h(x_{1}) - \ldots - h(x_{g}) + f(x_{1}, \ldots, x_{g}, x) \\ & = & \int_{X} \rho (x_{1}, \ldots, x_{g}, \cdot) h - \Big[ h(x_{1}) - G(x, x_{1}) \Big] - \ldots - \Big[ h(x_{g}) - G(x, x_{g}) \Big] \\ & - & \Big[ G(x, x_{1}) + \ldots + G(x, x_{g}) - f(x_{1}, \ldots, x_{g}, x) \Big] ~~, \end{eqnarray*} where $G$ is the Green's function of $X$. Because $h$ has only a logarithmic singularity it is allowed to differentiate under the integral sign. So the integral is smooth. The other summands are solutions of equations of the form $dd^{c} F = \sigma$ with a smooth $(1,1)$-form $\sigma$ on $X$ satisfying $\int_{X} \sigma = 0$ (in $x_{1}, \ldots, x_{g}$, respectively $x$). Since $dd^{c}$ is elliptic, these solutions exist as smooth functions and are unique up to constants. In particular, also the last summand must depend smoothly on $(x_{1}, \ldots, x_{g})$, even when some of the $x_{i}$ equal $x$. Note that the symmetry of the Green's function is used here essentially. \begin{center} $\Box$ \end{center} } \end{thm} {}~ \newline {\footnotesize {\bf Acknowledgement.} When doing this work, the author had fruitful discussions with U. Bunke (Berlin), who explained him much of the analytic part of the theory. He thanks him warmly.} \newpage {}~ \vspace*{-1.20truecm} \thispagestyle{myheadings} \markright{\rm ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~Jahnel} \small
\section{Introduction} The standard model\ of strong and electroweak\ interactions, though extremely successful in describing present day data, is known to be plagued by a number of shortcomings. There is no doubt that it is merely the low energy limit of a more profound underlying theory. In particular, the peculiar similarities between the quark and lepton sectors and their miraculous anomaly cancellations are indicative of a deeper interconnection between these two types of particles. Many models have been proposed which establish a closer link between quark and lepton degrees of freedom. They generally involve a new class of fields which carry both lepton and baryon number and mediate lepton-quark transitions. Such bosons can come in many combinations of the different quantum numbers and are generically called leptoquark s. They can, for example, emerge as composite scalar or vector states of techni-fermions \cite{etc} or preons \cite{comp}. Leptoquarks also arize naturally as gauge vectors or Higgs scalars in many grand unified models \cite{gut} or superstring-derived models \cite{ss}. In principle, $e^-p$ scattering provides the privileged reaction for discovering leptoquark s. Direct searches at HERA \cite{brw,ed} should be able to exclude leptoquark s below {\em ca} 300 GeV and a coupling to leptons and quarks above $0.03e$. The LEP-LHC combination could of course extend these limits much further and would also provide a powerful tool for discriminating different leptoquark\ types \cite{slava}. However, today the most stringent limits still originate from low-energy experiments \cite{sacha}. These bounds could be greatly improved at linear collider s of the next generation, though. In particular, the \mbox{$e^+e^-$}\ mode is very promising because it can abundantly pair-produce leptoquark s even if their couplings to leptons and quarks are tiny \cite{br}. In this paper we consider the production of single leptoquark s in \mbox{$e^-\gamma$}\ collisions, with laser backscattered photons. This reaction proceeds inevitably via the leptoquark-lepton-quark coupling, and may thus be suppressed if the latter is small. However, it may probe much higher leptoquark\ masses than the electron-positron annihilation \cite{br} or photon-gluon fusion \cite{ed} processes, which necessarily need to produce two leptoquark s. Moreover, since the standard model\ background s can easily be rendered harmless the data analysis should be exceedingly simple, in contrast to the electron-quark fusion reaction \cite{brw}. The angular distributions of the emerging leptoquark s have complicated patterns, which can be used advantageously to tell apart different types of leptoquark s. The study of scalar leptoquark\ production in \mbox{$e^-\gamma$}\ collisions was first discussed in Ref.~\cite{hp}. Similarly, vector leptoquark s were considered in Ref.~\cite{ce}. It was also shown \cite{nl} that even kinematically inaccessible scalar leptoquark s can also be probed this way if their couplings to fermions are large. For lighter scalar leptoquark s, it was pointed out \cite{bln} that substantial rates can be obtained even with Weizs\"acker-Williams photons in \mbox{$e^+e^-$}\ scattering. Because of the hadronic content of the photon, leptoquark s can also be probed via electron-quark fusion in \mbox{$e^-\gamma$}\ scattering \cite{eboli}. It was also shown that this resolved photon contribution may help determining leptoquark\ properties \cite{dg}. In the next section we introduce a model-independent framework for describing a large class of leptoquark s \cite{brw} and provide the analytical expressions of the integrated cross section s for producing the various types of leptoquark s. We then shortly review the generation and properties of the photon beam \cite{ginzburg}. After this, we discuss the leptoquark\ discovery potential of \mbox{$e^-\gamma$}\ scattering, which is conveniently summarized by Eq.~(\ref{osc}). Finally, in the last section we discuss the possibility of determining the nature of the discovered leptoquark s with the help of polarized beams and at hand of their angular distributions. \section{Cross sections} Because of the large number of possible leptoquark\ types, it is important to perform an analysis which is as model-independent as possible. Therefore, to describe the leptoquark-lepton-quark interactions we use the most general $SU(3)_c \otimes SU(2)_L \otimes U(1)_Y$ invariant effective lagrangian of lowest dimension which conserves lepton $(L)$ and baryon $(B)$ number. It can be separated into two parts, each involving either leptoquark s which carry no fermion number $F=3B+L=0$, or leptoquark s with fermion number $F=2$ \cite{brw}: \begin{eqnarray} {\cal L}_{F=0} \quad = & & \bigl( h_{2L} \bar{u}_R \ell_L + h_{2R} \bar{q}_L i \sigma_2 e_R \bigr) R_2 + \tilde{h}_{2L} \bar{d}_R \ell_L \tilde{R}_2 \\ & + & \bigl( h_{1L} \bar{q}_L \gamma^\mu \ell_L + h_{1R} \bar{d}_R \gamma^\mu e_R \bigr) U_{1 \mu} + \tilde{h}_{1R} \bar{u}_R \gamma^\mu e_R \tilde{U}_{1 \mu} \nonumber\\ & + & h_{3L} \bar{q}_L \mbox{\boldmath$\sigma$} \gamma^\mu \ell_L \mbox{\boldmath$U$}_{3 \mu} \nonumber\\ & + & \mbox{ h.c.} \nonumber\\ \nonumber\\ {\cal L}_{F=2} \quad = & & \bigl( g_{1L} \bar{q}^c_L i\sigma_2 \ell_L + g_{1R} \bar{u}^c_R e_R \bigr) S_1 + \tilde{g}_{1R} \bar{d}^c_R e_R \tilde{S}_1 \\ & + & g_{3L} \bar{q}^c_L i\sigma_2\mbox{\boldmath$\sigma$} \ell_L \mbox{\boldmath$S$}_3 \nonumber\\ & + & \bigl( g_{2L} \bar{d}^c_R \gamma^\mu \ell_L + g_{2R} \bar{q}^c_L \gamma^\mu e_R \bigr) V_{2 \mu} + \tilde{g}_{2L} \bar{u}^c_R \gamma^\mu \ell_L \tilde{V}_{2 \mu} \nonumber\\ & + & \mbox{ h.c.} \nonumber \end{eqnarray} where the \boldmath$\sigma$'s\unboldmath\ are Pauli matrices, while $q_L$ and $\ell_L$ are the $SU(2)_L$ quark and lepton doublets and $u_R$, $d_R$, $\ell_R$ are the corresponding singlets. The subscripts of the leptoquark s indicates the size of the $SU(2)_L$ representation they belong to. The $R$- and $S$-type leptoquark s are spacetime scalars, whereas the $U$ and $V$ are vectors. Family and colour indices are implicit. Since all these leptoquark s carry an electric charge, they must also couple to the photon. These interactions are described by the kinetic lagrangians for scalar and and vector bosons \cite{ed} \begin{eqnarray} {\cal L}_{J=0} & = & \sum_{\rm{scalars}} \quad \left(D_\mu\Phi\right)^{\dag} \left(D^\mu\Phi\right) - m^2 \Phi^{\dag}\Phi \\ {\cal L}_{J=1} & = & \sum_{\rm{vectors}} \quad -{1\over2} \left(D_\mu\Phi^\nu-D_\nu\Phi^\mu\right)^{\dag} \left(D^\mu\Phi_\nu-D^\nu\Phi_\mu\right) + m^2 \Phi_\mu^{\dag}\Phi^\nu \label{comp} \end{eqnarray} with the covariant derivative \begin{equation} D_\mu = \partial_\mu - ieQA_\mu \ , \end{equation} where $\Phi$ and $A$ are the leptoquark\ and photon fields, $m$ and $Q$ are the leptoquark\ mass and electromagnetic\ charge and $e$ is the electromagnetic\ coupling constant. This lagrangian describes the minimal vector boson coupling, typical of a composite leptoquark. If, however, the vector leptoquark s are gauge bosons, an extra Yang-Mills piece has to be added in order to maintain gauge invariance \cite{ed}: \begin{equation} {\cal L}_{G} = \sum_{\rm{vectors}} \quad -ie ~\Phi_\mu^{\dag}\Phi_\nu \left(\partial^\mu A^\nu - \partial^\nu A^\mu\right) \ .\label{ym} \end{equation} \begin{comment}{ If this piece is not included, tree-level unitarity is bound to be lost. This, however, is expected, since the effective lagrangian (\ref{comp}) is no longer valid at energies of the order of the compositeness scale. At those higher energies the terms which decouple at lower energies become relevant and the full (gauge) theory from which (\ref{comp}) was derived has to be considered. }\end{comment} Ignoring the resolved photon contributions, the following leptoquark\ reactions are possible to lowest order in \mbox{$e^-\gamma$}\ collisions: \begin{eqnarray*} F=0:~\left\{ \begin{array}{rl} ~J=0:~\left\{ \begin{array}{@{~}r@{~}c@{~}l} e^-_R ~ \gamma & \to & u_L ~ R_2^{-5/3} \\ && d_L ~ R_2^{-2/3} \\ \end{array} \qquad \begin{array}{@{~}r@{~}c@{~}l} e^-_L ~ \gamma & \to & u_R ~ R_2^{-5/3} \\ && d_R ~ \tilde R_2^{-2/3} \\ \end{array} \right. \\\\ ~J=1:~\left\{ \begin{array}{@{~}r@{~}c@{~}l} e^-_R ~ \gamma & \to & u_R ~ \tilde U_1^{-5/3} \\ && d_R ~ U_1^{-2/3} \\\\ \end{array} \qquad \begin{array}{@{~}r@{~}c@{~}l} e^-_L ~ \gamma & \to & u_L ~ U_3^{-5/3} \\ && d_L ~ U_1^{-2/3} \\ && d_L ~ U_3^{-2/3} \\ \end{array} \right. \end{array} \right. \label{f0} \end{eqnarray*} \begin{eqnarray*} F=2:~\left\{ \begin{array}{rl} ~J=0:~\left\{ \begin{array}{@{~}r@{~}c@{~}l} e^-_R ~ \gamma & \to & \bar u_R ~ S_1^{-1/3} \\ && \bar d_R ~ \tilde S_1^{-4/3} \\\\ \end{array} \qquad \begin{array}{@{~}r@{~}c@{~}l} e^-_L ~ \gamma & \to & \bar u_L ~ S_1^{-1/3} \\ && \bar u_L ~ S_3^{-1/3} \\ && \bar d_L ~ S_3^{-4/3} \\ \end{array} \right. \\\\ ~J=1:~\left\{ \begin{array}{@{~}r@{~}c@{~}l} e^-_R ~ \gamma & \to & \bar u_L ~ V_2^{-1/3} \\ && \bar d_L ~ V_2^{-4/3} \\ \end{array} \qquad \begin{array}{@{~}r@{~}c@{~}l} e^-_L ~ \gamma & \to & \bar u_R ~ \tilde V_2^{-1/3} \\ && \bar d_R ~ V_2^{-4/3} \\ \end{array} \right. \end{array} \right. \label{f2} \end{eqnarray*} The superscripts of the leptoquark s indicate their electromagnetic\ charge. The typical $s$-, $t$- and $u$-channel Feynman diagrams for the $F=0$ and $F=2$ leptoquark\ production are shown in Fig.~\ref{feyn}. The cross section s, though, do not depend explicitly on this quantum number. There are 24 different types of processes, depending on whether the produced leptoquark\ is a scalar, vector or gauge boson, whether it couples to right- or left-handed leptons and what is its electromagnetic\ charge ($Q=-1/3,-2/3,-4/3,-5/3$). The differential cross section s are lengthy and we do not report them here. We agree with the unpolarized expressions reported in Refs~\cite{hp,ce,nl}\footnote{ The colour factor is not explicitly stated in Refs~\cite{hp,nl}. }, though. Let us define \begin{equation} x = {m^2 \over s} \ ,\label{x} \end{equation} where $m$ is the leptoquark\ mass and $s$ is the centre of mass\ energy squared. We also use the generic leptoquark\ coupling $\lambda$ to the electrons and quarks, and denote the electron and photon polarizations $P_e$ and $P_\gamma$. We find for the integrated scalar cross section s \begin{eqnarray} \makebox[0cm][l]{\hskip-5mm$\displaystyle \sigma(J=0) \quad = \quad {3\pi\alpha^2\over2s} \quad \left({\lambda\over e}\right)^2 \quad {1 \pm P_e \over 2} \quad \times$} \label{s0}\\ & \biggl\{\quad & \left(-\left(3+4Q\right) + \left(7+8Q+8Q^2\right)x\right) \quad (1-x) \nonumber\\ && +4Q\left( Q - \left(2+Q\right)x \right) \quad x \ln x \nonumber\\ && -2\left(1+Q\right)^2\left( 1 - 2x + 2x^2 \right) \quad \ln{m_q^2/s\over\left(1-x\right)^2} \nonumber\\ & \pm~P_\gamma~\Bigl[~ & \left(-\left(7+12Q+4Q^2\right) + 3x\right) \quad (1-x) \nonumber\\ && +4Q^2 \quad x \ln x \nonumber\\ && -2\left(1+Q\right)^2\left( 1 - 2x \right) \quad \ln{m_q^2/s\over\left(1-x\right)^2} \quad \Bigr] \quad \biggr\} \nonumber~. \end{eqnarray} The unpolarized part ($P_e=P_\gamma$=0) of this expression agrees with the result reported in Ref.~\cite{bln}, up to the colour factor, which we include explicitly here. The vector cross section s are \begin{eqnarray} \makebox[0cm][l]{\hskip-5mm$\displaystyle \sigma(J=1) \quad = \quad {3\pi\alpha^2\over8m^2} \quad \left({\lambda\over e}\right)^2 \quad {1 \pm P_e \over 2} \quad \times$} \label{c0}\\ & \biggl\{\quad & \left(Q^2 + \left(8-16Q-Q^2\right)x + 8\left(7+8Q+8Q^2\right)x^2\right) \quad (1-x) \nonumber\\ && -4Q\left( Q + \left(8+3Q\right)x - 8Qx^2 + 8\left(2+Q\right)x^3 \right) \quad \ln x \nonumber\\ && -16\left(1+Q\right)^2\left( 1 - 2x + 2x^2 \right) \quad x \ln{m_q^2/s\over\left(1-x\right)^2} \nonumber\\ & \pm~P_\gamma~\Bigl[~ & \left(- 3Q^2 + \left(40+48Q+63Q^2\right)x + 24x^2\right) \quad (1-x) \nonumber\\ && -4Q\left( -3Q + 8\left(3+Q\right)x \right) \quad x \ln x \nonumber\\ && +16\left(1+Q\right)^2\left( 1 - 2x \right) \quad x \ln{m_q^2/s\over\left(1-x\right)^2} \quad \Bigr] \quad \biggr\} \nonumber~. \end{eqnarray} Finally, if the vector is a gauge field, we have \begin{eqnarray} \makebox[0cm][l]{\hskip-5mm$\displaystyle \sigma^{\rm YM}(J=1) \quad = \quad {3\pi\alpha^2\over m^2} \quad \left({\lambda\over e}\right)^2 \quad {1 \pm P_e \over 2} \quad \times$} \label{v0}\\ & \biggl\{\quad & \left(4Q^2 + \left(1-4Q\right)x + \left(7+8Q+8Q^2\right)x^2\right) \quad (1-x) \nonumber\\ && -4Q\left( 2 - Qx + \left(2+Q\right)x^2 \right) \quad x \ln x \nonumber\\ && -2\left(1+Q\right)^2\left( 1 - 2x + 2x^2 \right) \quad x \ln{m_q^2/s\over\left(1-x\right)^2} \nonumber\\ & \pm~P_\gamma~\Bigl[~ & \left(\left(5+4Q+12Q^2\right) + 3x\right) x \quad (1-x) \nonumber\\ && +4Q\left( Q - \left(4+Q\right)x \right) \quad x \ln x \nonumber\\ && +2\left(1+Q\right)^2\left( 1 - 2x \right) \quad x \ln{m_q^2/s\over\left(1-x\right)^2} \quad \Bigr] \quad \biggr\} \nonumber~. \end{eqnarray} The electron and quark masses have been set equal to zero everywhere to derive Eqs~(\ref{s0}-\ref{v0}), except in the squared $u$-channel matrix elements. It is essential to perform the calculation of these terms with a finite quark mass, because it regulates the singularity which occurs when a leptoquark\ is emitted in the direction of the incoming electron. The approximation $\ln m_q^2/s/(1-x)^2$ we have written for this $u$-channel logarithm stays more than accurate up to within a few quark masses from the threshold. It is straightforward to compute the full expression \cite{bln}. Moreover, the integration in the backward direction introduces non-vanishing contributions of the order ${\cal O}(m_q^2/m_q^2)$ to the non-logarithmic parts of the cross section s (\ref{s0}-\ref{v0})\footnote{ I am very much indebted to David London and H\'el\`ene Nadeau for pointing this out to me. }. The threshold behaviour of the cross section s, around $x=1$, mainly depends on the $u$-channel singularity and on the electron and photon relative polarizations: \begin{equation} x=1: \left\{ \quad \begin{array}{l} \displaystyle{\sigma(J=0) \quad \propto \quad ~(1+Q)^2 \quad (1 \mp P_\gamma) (1 \pm P_e) } \\\\ \displaystyle{\sigma(J=1) \quad \propto \quad 2(1+Q)^2 \quad (1 \pm P_\gamma) (1 \pm P_e) } \end{array} \right. \label{thresh} \end{equation} The scalar cross section\ quickly drops to zero for like-sign initial state electrons and photons polarizations, whereas the vector cross section\ is suppressed when these polarizations have opposite signs. For equal couplings, the vector threshold cross section s are twice as intense as the scalar ones. In the asymptotic region, for $x=0$, the $J=0$ and $J=1$ leptoquark s also display very different behaviours. Whereas the scalar cross section s decrease like $1/s$, the vector cross section s eventually increase like $\ln s$. If the vectors are gauge fields, though, their cross section s saturate for large values of $s$. \section{Photon Beams} The cross section s (\ref{s0}-\ref{v0}) still have to be folded with a realistic Compton backscattered photon spectrum \cite{ginzburg}: \def\,{\rm d}{\,{\rm d}} \begin{equation} \sigma(s) = \int_{x_{\rm min}}^{x_{\rm max}} \,{\rm d} x {\,{\rm d} n(x) \over \,{\rm d} x} \sigma(xs) \ ,\label{fold} \end{equation} where the probability density of a photon to have the energy fraction $x = E_\gamma/E_{\rm beam}$ is given by \begin{eqnarray} {\,{\rm d} n(x) \over \,{\rm d} x} = \frac{1}{\cal N} & \Biggl\{ & 1 - x + \frac{1}{1 - x} - \frac{4 x }{z ( 1 - x) } + \frac{4 x^2 } {z^2 (1 - x)^2} \nonumber \\ & + & P_{\rm beam} P_{\rm laser} \frac{x (2 - x)}{1 - x} \left[ \frac{2 x} {z (1 - x)} - 1 \right] \Biggr\} \ , \label{spec} \end{eqnarray} where \begin{eqnarray} {\cal N} & = & \frac{z^3 + 18 z^2 + 24 z + 8}{2 z (z + 1)^2 } + \left( 1 - \frac{4}{z} - \frac{8}{z^2} \right) \ln (1 + z) \\ & + & P_{\rm beam} P_{\rm laser} \left[ 2 - \frac{z^2}{(z + 1)^2} - \left( 1 + \frac{2}{z} \right) \ln (1 + z) \right] \ ,\nonumber \end{eqnarray} \begin{equation} z = \frac{4 E_{\rm beam} E_{\rm laser} }{m_e^2} \ , \end{equation} $m_e$ being the mass of the electron, By design, the energy fraction $x$ of the photons is limited from above to \begin{equation} x_{\rm max} = \frac{ 2+2\sqrt{2}}{ 3+2\sqrt{2}} \approx 0.8284 \label{xmax} \end{equation} in order to prevent electron-positron pair-production from photon rescattering. In practice, it is also limited from below, because only the harder photons are produced at a small angle with respect to\ the beam-pipe according to \begin{equation} \theta_\gamma (x) \simeq {m_e \over E_{\rm beam}} \sqrt{{z \over x}-z-1} \ . \end{equation} The softer photons are emitted at such large angles that they are bound to miss the opposite highly collimated electron beam. Assuming a conversion distance of {\em ca} 5 cm and a beam size of 500 nm diameter, the lower bound we adopt for the photon energy fraction is \begin{equation} x_{\rm min} = .5 \ . \label{xmin} \end{equation} Of course, if the beams are very flat, this cut-off will be somewhat softened out. However, in the threshold region is has no effect. When it sets in, at $s=m^2/x_{\rm min}=2m^2$, it is visible on the plots of Fig.~\ref{eny} as a slight kink in the slopes. The polarization of the backscattered photons is given by \begin{equation} P_\gamma (x) = \frac{P_{\rm laser} \zeta (2 - 2 x + x^2) + P_{\rm beam} x ( 1 + \zeta^2)} { (1 -x ) (2 - 2 x + x^2) - 4 x ( z - z x - x)/z^2 - P_{\rm beam} P_{\rm laser} \zeta x ( 2 - x) } \ , \label{polar} \end{equation} with \begin{equation} \zeta = 1 - x ( 1 + 1/z) \ . \label{e103} \end{equation} All these features of the Compton backscattered photon beams are displayed in Fig.~\ref{photon}. In particular, it appears that if the signs of the polarizations of the laser and the Compton scattered electron beams are chosen to be opposite, we obtain the most intense, hard and highly polarized photon beam. We adopt this choice throughout the rest of this paper. \section{Leptoquark Discovery Limits} To present our results, we have chosen to work with a quark mass in the $u$-channel propagator $m_q=10$ MeV. Since the dependence of the cross section s on this mass is only logarithmic, the error does not exceed a few \%\ and is mainly confined to the immediate threshold region. We also assumed both the unaltered and the Compton scattered electron beams to be 90\%\ polarized. This should be fairly easy to realize at a linear collider\ of the next generation. In Fig.~\ref{eny} we have displayed the behaviours of the cross section s (\ref{s0}-\ref{v0}) as functions of the collider centre of mass\ energy. For the purpose of these plots, we have set the leptoquark-electron-quark coupling equal to the electromagnetic\ coupling constant $\lambda=e$. In general, leptoquark s which are produced in the reactions (\ref{s0}-\ref{v0}), will decay into a charged lepton and a jet with a substantial branching ratio. If the leptoquark s are bound to a single generation, the decay lepton is an electron. Around threshold, the $u$-channel pole is dominant, so the quarks and leptoquark s are mostly produced at very small angles from the beampipe. Since the leptoquark s are almost at rest, most of the electron and quark into which they decay are emitted at large angles. The leptoquark\ signature is thus a low transverse momentum\ (or even invisible) jet, and a high transverse momentum\ electron and jet pair whose invariant mass is closely centered around the leptoquark\ mass. Away from threshold, the $u$-channel is no longer dominant, and the leptoquark\ signature becomes a high transverse momentum\ electron and pair of jets, where the electron and one of the jets have a combined invariant mass close to the leptoquark\ mass. If all jet pairs with an invariant mass around the $Z^0$ mass are cut out, there is no $Z^0$ Compton scattering background. The next order background which then subsists is the quark photoproduction process $\mbox{$e^-\gamma$} \to e^-q\bar q$. This background, though, will not show any peak in the electron-jet invariant mass distribution at the leptoquark\ mass, and can moreover be drastically reduced by requiring a minimum transverse momentum\ for the electron. If one allows for leptoquark s which couple equally well to different generations, the final state can contain a muon instead of an electron. In this case, of course, there is no standard model\ background at all. To estimate the leptoquark\ discovery potential of \mbox{$e^-\gamma$}\ collisions, we have plotted in Fig.~\ref{lim} the boundary in the $(m_{LQ},\lambda/e)$ plane, below which the cross section\ $\sigma(J=1,Q=-5/3)$ in Eq.~(\ref{v0}) yields less than 10 events. For this we consider four different collider energies and assume 10 fb$^{-1}$ of accumulated luminosity. In general, as can be inferred from Eq.~(\ref{thresh}), these curves are closely osculated by the relation \begin{equation} {\lambda \over e} \quad = \quad 0.03 \quad {m/\mbox{TeV} \over \sqrt{{\cal L}/{\rm fb}^{-1}}} \quad \sqrt{{n \over (J+1) (1+Q)^2 }} \qquad \left( m \le .8\sqrt{s_{ee}} \right) \ ,\label{osc} \end{equation} where $\lambda$ is the leptoquark's coupling to leptons and quarks, $m$ its mass, $Q$ its charge, $J$ its spin, $n$ is the required number of events and $\cal L$ is the available luminosity. This scaling relation provides a convenient means to gauge the leptoquark\ discovery potential of \mbox{$e^-\gamma$}\ scattering. It is valid for leptoquark\ masses short off 80\%\ of the collider energy and assumes the electron beams to be 90\%\ polarized ($|P_e|=.9$) while the chirality of the fully polarized photon beam is chosen such as to enhance the threshold cross section\ ({\em cf.} Eq.~(\ref{thresh})). In comparison, the best bounds on the leptoquark\ couplings which have been derived indirectly from low energy data \cite{sacha} are no better than \begin{equation} {\lambda \over e} \quad \ge \quad m/\mbox{TeV} \ , \end{equation} for leptoquark\ interactions involving only the first generation. Similar bounds on couplings involving higher generations are even poorer. \section{Leptoquark-Type Discrimination} If a leptoquark\ is discovered someday, it is interesting to determine its nature. In principle, \mbox{$e^-\gamma$}\ scattering may discriminate between the 24 combinations of the quantum numbers $J$, $Q$ and $P$, where the latter is the chirality of the electron to which the leptoquark\ couples. For the case $J=1$ we make the distinction between gauge and non-gauge leptoquark s. It is of course trivial to determine $P$ by switching the polarization of the electron beam. Similarly, it is almost as easy to distinguish scalars from vectors. Indeed, as can be inferred from Eqs~(\ref{thresh}) and from Fig.~\ref{eny}, all threshold cross section s are very sensitive to the relative electron and photon polarizations. Since this effect works in opposite directions for $J=0$ and $J=1$, a simple photon polarization flip upon discovery should suffice to determine the spin of the discovered leptoquark s. Determining the other quantum numbers, unfortunately, is not as easy. In practice this task is not facilitated by the fact all the different models involving leptoquark s predict very different values for their couplings to leptons and quarks, if at all. Basically, we have no idea what to expect. Fig.~\ref{eny} indicates that it may be possible to discriminate some of the different vector leptoquark s, by combining the information gathered from polarization flips and an energy scan. It will, however, be exceedingly difficult to differentiate for instance the different scalar leptoquark s from eachother using the information from total cross section s alone. Much more discriminating power can be obtained from the differential distributions, though. We do not report their long analytical forms here. As it turns out, the interferences between the different channels result into rather complex angular dependences of the cross section s. There are even radiation zeros for the reactions involving scalar and Yang-Mills leptoquark s of charges -1/3 and -2/3. Note that in the cross-channel reaction $e^-q \to LQ\gamma$ these radiation zeros occur for the scalar and Yang-Mills leptoquark s of charges -4/3 and -5/3 \cite{slava}. These very salient features could well be observed away from threshold, where the $u$-channel pole is no longer so dominant. Of course, the convolution with the photon energy and polarization spectra washes out to some extent these prominent features. But, as can be gathered from Fig.~\ref{ang}, even so the angular distributions of the different leptoquark s retain their distinctive characteristics. In this figure, we have considered 400 GeV leptoquark s produced at a 1 TeV collider. To roughly estimate the \mbox{$e^-\gamma$}\ potential for discriminating the different types of leptoquark s, we compare these differential distributions with a Kolmogorov-Smirnov test. Say $a$ and $b$ are two different kinds of leptoquark s and $a$ is the one which is observed. The probability that $b$ could be mistaken for $a$ is related to the statistic \begin{equation} D~=~\sqrt{N_a} ~\sup \left| {1\over\sigma_a}\int^\theta\,{\rm d}\theta{\,{\rm d}\sigma_a\over\,{\rm d}\theta} - {1\over\sigma_b}\int^\theta\,{\rm d}\theta{\,{\rm d}\sigma_b\over\,{\rm d}\theta} \right| \label{eks}\ , \end{equation} where $\theta$ is the polar angle of the leptoquark\ and $N_a={\cal L}\sigma_a$ is the number of observed events. Focusing on the case of Fig.~\ref{ang}, where a 400 GeV leptoquark\ is studied at a 1 TeV collider, Tables~\ref{ts} and \ref{tv} summarize the minimal values of $$ {\lambda \over e} \quad \sqrt{{\cal L}/{\rm fb}^{-1}} $$ needed to tell apart two different types of leptoquark s with 99.9\%\ confidence, {\em i.e.}, setting $D=1.95$ in Eq.~(\ref{eks}). Assuming each combination of electron and photon polarizations has accumulated 50 fb$^{-1}$ of data, some leptoquark\ types have so different angular distributions that a coupling as small as $\lambda=.0085e$ is quite sufficient to tell them apart. Others need as much as $\lambda=.31e$. Obviously, these numbers are only valid for this particular choice of leptoquark\ mass and collider energy, and could be improved with a more sophisticated analysis. Nevertheless, they should be indicative of what resolving power an \mbox{$e^-\gamma$}\ scattering experiment can achieve. \section{Conclusions} We have studied leptoquark\ production in the \mbox{$e^-\gamma$}\ mode of a linear collider\ of the next generation. To perform this analysis, we have considered all types of scalar and vector leptoquark s, whose interactions with leptons and quarks conserve lepton and baryon number and are invariant under the standard model\ $SU(3)_c \otimes SU(2)_L \otimes U(1)_Y$ gauge group. For the vectors we distinguish between leptoquark s which couple minimally to photons and Yang-Mills fields. The advantage of an \mbox{$e^-\gamma$}\ experiment over, {\em e.g.}, \mbox{$e^+e^-$}\ annihilation is that the leptoquark s need not be produced in pairs. Hence higher masses can be explored. The disadvantage is that the \mbox{$e^-\gamma$}\ reactions proceed only via the {\em a priori} unknown leptoquark-lepton-quark couplings. The standard model\ backgrounds can be reduced to almost zero by simple kinematical cuts, and the discovery potential is conveniently summarized by the scaling relation Eq.~(\ref{osc}). The spin of the leptoquark s can easily be determined at threshold by inverting the polarization of the photon beam. Moreover, each leptoquark\ type displays very characteristic angular distributions, some of which having even radiation zeros. Therefore, the prospects for discriminating different leptoquark s of the same spin can be greatly enhanced by studying angular correlations. All our results have been obtained with a realistic electron beam polarization and Compton backscattered photon spectra. \begin{ack} It is a pleasure to thank Stan Brodsky and Clem Heusch for their hospitality at SLAC and UCSC, where this project was initiated. Many fruitful discussions with Sacha Davidson, Paul Frampton, Slava Ilyin, David London and H\'el\`ene Nadeau are gratefully acknowledged. I am also particularly indebted to David London for pointing out Refs~\cite{hp,ce,nl,bln,eboli,dg} to me and to H\'el\`ene Nadeau for helping me eliminate an error from my calculations. \end{ack}
\section{Introduction} Rational vertex operator algebras, which play a fundamental role in rational conformal field theory (see [BPZ] and [MS]), single out an important class of vertex operator algebras. Most vertex operator algebras which have been studied so far are rational vertex operator algebras. Familiar examples include the moonshine module $V^{\natural}$ ([B], [FLM], [D2]), the vertex operator algebras $V_L$ associated with positive definite even lattices $L$ ([B], [FLM], [D1]), the vertex operator algebras $L(l,0)$ associated with integrable representations of affine Lie algebras [FZ] and the vertex operator algebras $L(c_{p,q},0)$ associated with irreducible highest weight representations for the discrete series of the Virasoro algebra ([DMZ] and [W]). A rational vertex operator algebra as studied in this paper is a vertex operator algebra such that any {\em admissible} module is a direct sum of simple ordinary modules (see Section 2). It is natural to ask if such complete reducibility holds for an arbitrary weak module (defined in Section 2). A rational vertex operator algebra with this property is called a {\em regular} vertex operator algebra. One motivation for studying such vertex operator algebras arises in trying to understand the appearance of negative fusion rules (which are computed by the Verlinde formula) for vertex operator algebras $L(l,0)$ for certain rational $l$ (cf. [KS] and [MW]). In this paper we give several sufficient conditions under which a rational vertex operator algebra is regular. We prove that the rational vertex operator algebras $V^{\natural},$ $L(l,0)$ for positive integers $l,$ $L(c_{p,q},0)$ and $V_L$ for positive definite even lattices $L$ are regular. Our result for $L(l,0)$ implies that any restricted integrable module of level $l$ for the corresponding affine Lie algebra is a direct sum of irreducible highest weight integrable modules. This result is expected to be useful in comparing the construction of tensor product of modules for $L(l,0)$ in [F] based on Kazhdan-Lusztig's approach [KL] with the construction of tensor product of modules [HL] in this special case. We should remark that $V_L$ in general is a vertex algebra in the sense of [DL] if $L$ is not positive definite. In this case we establish the complete reducibility of any weak module. Since the definition of vertex operator algebra is by now well-known, we do not define vertex operator algebra in this paper. We refer the reader to [FLM] and [FHL] for their elementary properties. The reader can find the details of the constructions of $V^{\natural}$ and $V_L$ in [FLM], and $L(l,0)$ and $L(c_{p,q},0)$ in [DMZ], [DL], [FLM], [FZ], [L1] and [W]. The paper is organized as follows: In Section 2, after defining the notion of weak module for a vertex operator algebra and the definition of rational vertex operator algebra, we discuss the rational vertex operator algebras $V^{\natural},$ $V_L,$ $L(l,0)$ and $L(c_{p,q},0).$ Section 3 is devoted to regular vertex operator algebras. We begin this section with the definition of regular vertex operator algebra. We show that the tensor product of regular vertex operator algebras is also regular and that a rational vertex operator algebra is regular under either of the assumptions (i) it contains a regular vertex operator subalgebra, or (ii) any weak module contains a simple ordinary module. These results are then used to prove that $V^{\natural},$ $V_L$ ($L$ is positive definite), $L(l,0)$ and $L(c_{p,q},0)$ are regular. We also discuss the complete reducibility of weak $V_L$-modules for an arbitrary even lattice $L.$ Based on these results, we conjecture that {\em any} rational vertex operator algebra is regular. We thank Yi-Zhi Huang for pointing out a mistake in a prior version of this paper. \section{Rational vertex operator algebras} Let $(V,Y,{\bf 1},\omega)$ be a vertex operator algebra (cf. [B], [FHL] and [FLM]). A {\em weak module} $M$ for $V$ is a vector space equipped with a linear map $$\begin{array}{l} V\to (\mbox{End}\,M)[[z^{-1},z]]\label{map}\\ v\mapsto\displaystyle{ Y_M(v,z)=\sum_{n\in\Bbb Z}v_nz^{-n-1}\ \ \ (v_n\in \mbox{End}\,M)} \end{array}$$ (where for any vector space $W,$ we define $W[[z^{-1},z]]$ to be the vector space of $W$-valued formal series in $z$) satisfying the following conditions for $u,v\in V$, $w\in M$: \begin{eqnarray} & &Y_M(v,z)=\sum_{n\in \Bbb Z}v_nz^{-n-1}\ \ \ \ \mbox{for}\ \ v\in V;\label{1/2}\\ & &v_nw=0\ \ \ \mbox{for}\ \ \ n\in \Bbb Z \ \ \mbox{sufficiently\ large};\label{vlw0}\\ & &Y_M({\bf 1},z)=1;\label{vacuum} \end{eqnarray} \begin{equation}\label{jacobi} \begin{array}{c} \displaystyle{z^{-1}_0\delta\left(\frac{z_1-z_2}{z_0}\right) Y_M(u,z_1)Y_M(v,z_2)-z^{-1}_0\delta\left(\frac{z_2-z_1}{-z_0}\right) Y_M(v,z_2)Y_M(u,z_1)}\\ \displaystyle{=z_2^{-1}\delta\left(\frac{z_1-z_0}{z_2}\right) Y_M(Y(u,z_0)v,z_2)}. \end{array} \end{equation} \begin{equation}\label{vir} [L(m),L(n)]=(m-n)L(m+n)+\frac{1}{12}(m^3-m)\delta_{m+n,0}(\mbox{rank}\,V) \end{equation} for $m, n\in {\Bbb Z},$ where \begin{eqnarray} & &L(n)=\omega_{n+1}\ \ \ \mbox{for}\ \ \ n\in{\Bbb Z}, \ \ \ \mbox{i.e.},\ \ \ Y_M(\omega,z)=\sum_{n\in{\Bbb Z}}L(n)z^{-n-2};\nonumber\\ & &\frac{d}{dz}Y_M(v,z)=Y_M(L(-1)v,z).\label{6.72} \end{eqnarray} This completes the definition. We denote this module by $(M,Y_M)$ (or briefly by $M$). \begin{de}\label{r2.2} An (ordinary) $V$-module is a weak $V$-module which carries a $\Bbb C$-grading $$M=\coprod_{\lambda \in{\Bbb C}}M_{\lambda} $$ such that $\dim M_{\lambda }$ is finite and $M_{\lambda +n}=0$ for fixed $\lambda $ and $n\in {\Bbb Z}$ small enough. Moreover one requires that $M_{\lambda }$ is the $\lambda $-eigenspace for $L(0):$ $$L(0)w=\lambda w=(\mbox{wt}\,w)w, \ \ \ w\in M_{\lambda }.$$ \end{de} This definition is weaker than that of [FLM], for example, where the grading on $M$ is taken to be rational. The extra flexibility attained by allowing $\Bbb C$-gradings is important $-$ see for example [DLM1] and [Z]. We observe some redundancy in the definition of weak module: \begin{lem}\label{r2.1} Relations (\ref{vir}) and (\ref{6.72}) in the definition of weak module are consequences of (\ref{1/2})-(\ref{jacobi}). \end{lem} {\bf Proof. } To establish (\ref{6.72}) note that $L(-1)u=L(-1)u_{-1}{\bf 1}=u_{-2}{\bf 1}$ for $u\in V.$ Then \begin{eqnarray} & &\ \ \ Y_{M}(L(-1)u,z_{2})\nonumber\\ & &=Y_{M}(u_{-2}{\bf 1},z_{2})\nonumber\\ & &={\rm Res}_{z_{0}}z_{0}^{-2}Y_{M}(Y(u,z_{0}){\bf 1},z_{2})\nonumber\\ & &={\rm Res}_{z_{0}}{\rm Res}_{z_{1}}z_{0}^{-2} \left(z^{-1}_0\delta\left(\frac{z_1-z_2}{z_0}\right) Y_M(u,z_1)Y_M({\bf 1},z_2)\right.\nonumber\\ & &\ \ \ \ \ \ \left.-z^{-1}_0\delta\left(\frac{z_2-z_1}{-z_0}\right) Y_M({\bf 1},z_2)Y_M(u,z_1)\right)\nonumber\\ & &={\rm Res}_{z_{0}}{\rm Res}_{z_{1}}z_{0}^{-2}z_2^{-1}\delta\left(\frac{z_1-z_0}{z_2}\right) Y_{M}(u,z_{1})\nonumber\\ & &={\rm Res}_{z_{0}}{\rm Res}_{z_{1}}z_{0}^{-2}z_1^{-1}\delta\left(\frac{z_2+z_0}{z_1}\right) Y_{M}(u,z_{2}+z_{0})\nonumber\\ & &={\rm Res}_{z_{0}}z_{0}^{-2}Y_{M}(u,z_{2}+z_{0})\nonumber\\ & &={\rm Res}_{z_{0}}z_{0}^{-2}e^{z_0\frac{d}{d z_2}}Y_{M}(u,z_{2})\nonumber\\ & &={d\over dz_{2}}Y_{M}(u,z_{2}). \end{eqnarray} This establishes (\ref{6.72}), and together with (\ref{jacobi}) and $Y(\omega,z_0)\omega=\frac{1}{2}({\rm rank}V)z_0^{-4} +2\omega z_0^{-2}+L(-1)\omega z_0^{-1}+$ regular terms we can easily deduce (\ref{vir}). \ \ \ $\Box$ Thus we may just use (\ref{1/2})-(\ref{jacobi}) as the axioms for a weak $V$-module. \begin{de}\label{d2.2} An {\em admissible} $V$-module is a weak $V$-module $M$ which carries a ${\Bbb Z}_{+}$-grading $$M=\coprod_{n\in {\Bbb Z}_{+}}M(n)$$ ($\Bbb Z_+$ is the set all nonnegative integers) satisfying the following condition: if $r, m\in {\Bbb Z} ,n\in {\Bbb Z}_{+}$ and $a\in V_{r}$ then \begin{eqnarray} a_{m}M(n)\subseteq M(r+n-m-1).\label{2.7} \end{eqnarray} We call an admissible $V$-module $M$ {\em simple} in case $0$ and $M$ are the only $\Bbb Z_+$-graded submodules. $V$ is called {\em rational} if every admissible $V$-module is a direct sum of simple admissible $V$-modules. That is, we have complete reducibility of admissible $V$-modules. \end{de} \begin{rem}\label{r2.4} (i) Note that any ordinary $V$-module is admissible. (ii) It is proved in [DLM1] that if $V$ is rational then conversely, every simple admissible $V$-module is an ordinary module. Moreover $V$ has only a finite number of inequivalent simple modules. (iii) Zhu's definition of rational vertex operator algebra $V$ is as follows [Z]: (a) all admissible $V$-module are completely reducible, (b) each simple admissible $V$-module is an ordinary $V$-module, (c) $V$ only has finitely many inequivalent simple modules. Thanks to (ii), Zhu's definition of rational thus coincides with our own. \end{rem} We next introduce a certain category $\cal{O}$ of admissible $V$-modules in analogy with the well-known category $\cal O$ of Bernstein-Gelfand-Gelfand. First some notation: for any weak $V$-module $M$ we set for $h\in \Bbb C:$ $$M_{h}=\{m\in M| (L(0)-h)^km=0\ {\rm for\ some} \ k\in\Bbb Z_{+}\}.$$ So $M_h$ is a {\em generalized} eigenspace for $L(0),$ and in particular $M_h$ is the $h$-eigenspace for $L(0)$ if $L(0)$ is a semisimple operator. Now define $\cal{O}$ to be the category of weak $V$-modules $M$ satisfying the following two conditions: (1) $L(0)$ is locally finite in the sense that if $m\in M$ then there is a finite-dimensional $L(0)$-stable subspace of $M$ which contains $m.$ (2) There are $h_{1}, \cdots, h_{k}\in\Bbb C$ such that $$M=\oplus _{i=1}^{k}\oplus _{n\in {\Bbb Z}_{+}}M_{n+h_{i}}.$$ These are the {\em objects} of $\cal O.$ Morphisms may be taken to be $V$-module homomorphisms, though we will not make use of them in the sequel. \begin{rem}\label{r2.5} (i) Any weak $V$-module which belongs to $\cal O$ is necessarily admissible: a $\Bbb Z_+$-grading obtains by defining $M(n)=\oplus_{i=1}^kM_{h_i+n}.$ Condition (\ref{2.7}) follows in the usual way. (ii) Suppose that $M$ is a weak $V$-module and that $W$ is a weak $V$-submodule of $M.$ Then $M$ lies in $\cal O$ if, and only if, both $W$ and $M/W$ lie in $\cal O.$ (iii) If $V$ is rational, any weak $V$-module in $\cal{O}$ is a direct sum of simple $V$-modules (use Remark \ref{r2.4} (ii)). \end{rem} Next we briefly discuss some familiar examples of rational vertex operator algebras. The reader is referred to the references for notation and the details of the constructions. (1) Let $L$ be an even lattice and $V_{L}$ the corresponding vertex algebra (see [B], [DL] and [FLM]). It is proved in [D1] that if $L$ is positive definite then $V_{L}$ is rational and its simple modules are parametrized by $L'/L$ where $L'$ is the dual lattice of $L.$ (2) Let ${\frak g}$ be a finite-dimensional simple Lie algebra with a Cartan subalgebra ${\frak h}$ and $\hat \frak g=\Bbb C[t,t^{-1}]\otimes \frak g\oplus\Bbb C c$ the corresponding affine Lie algebra. Fix a positive integer $l.$ Then any $\lambda\in \frak h^*$ can be viewed as a linear form on $\Bbb C c\oplus \frak h\subset \hat \frak g$ by sending $c$ to $l.$ Let us denote the corresponding irreducible highest weight module for $\hat\frak g$ by $L(l,\lambda).$ Then $L(\ell,0)$ is a rational vertex operator algebra ([DL], [FZ], [L1]). (3) Let $c$ and $h$ be two complex numbers and let $L(c,h)$ be the lowest weight irreducible module for the Virasoro algebra with central charge $c$ and lowest weight $h$. Then $L(c,0)$ has a natural vertex operator algebra structure (cf. [FZ]). Moreover, $L(\ell,0)$ is rational if, and only if, $c=c_{p,q}=1-\frac{6(p-q)^{2}}{pq}$ for $p,q\in \{2,3,4,\cdots\}$ and $p, q$ are relatively prime (see [DMZ] and [W]). (4) Let $V^{\natural}$ be the moonshine module vertex operator algebra constructed by Frenkel, Meurman and Lepowsky [FLM] (see also [B]). It is established in [D2] that $V^{\natural}$ is {\em holomorphic} in the sense that $V^{\natural}$ is rational and the only simple module is $V^{\natural}$ itself. (5) Let $V^{1},\cdots, V^{k}$ be vertex operator algebras. Then $V=\otimes_{i=1}^{k}V^{i}$ is a vertex operator algebra of rank $\sum_{i=1}^{k}{\rm rank} V^{i}$ and any simple $V$-module $M$ is isomorphic to a tensor product module $M^{1}\otimes \cdots \otimes M^{k}$ for some simple $V^{i}$-module $M^i$ [FHL]. Furthermore, $V^{1}\otimes V^{2}\otimes \cdots\otimes V^{k}$ is rational if, and only if, each $V^i$ is rational [DMZ]. (6) Let $V^{1}, \cdots, V^{k}$ be vertex operator algebras of the same rank. Then $\oplus_{i=1}^{k}V^{k}$ is a vertex operator algebra [FHL]. It is clear that $\oplus_{i=1}^{k}V^{i}$ is rational if each $V^{i}$ is rational. The vacuum space of the resulting vertex operator algebra is not one-dimensional, and we will not consider this particular example further in this paper. \section{Regular vertex operator algebras} In Section 2 we made use of the complete reducibility of admissible modules in order to define rational vertex operator algebras. The study of complete reducibility of an arbitrary weak module for a rational vertex operator algebra leads us to the following notion of regular vertex operator algebra: \begin{de}\label{d2.4} A vertex operator algebra $V$ is said to be {\em regular} if any weak $V$-module $M$ is a direct sum of simple ordinary $V$-modules. \end{de} \begin{rem}\label{r3.2} A regular vertex operator algebra $V$ is necessarily rational. Indeed if $M$ is a weak $V$-module then, being a direct sum of ordinary simple $V$-module, it is admissible (Remark \ref{r2.4} (i)) and, for the same reason, a direct sum of simple admissibles. \end{rem} The main result of the present paper is to show that the rational vertex operator algebras in examples (1)-(4) of Section 2 are each regular. First we have some general results which will be useful later. \begin{prop}\label{p2.5} Let $V^{1}, \cdots, V^{k}$ be regular vertex operator algebras. Then $V=V^{1}\otimes V^{2}\otimes \cdots \otimes V^{k}$ is regular. \end{prop} {\bf Proof.} Let $M$ be a weak $V$-module. For each $1\le i\le k$, we may regard $V^{i}$ as a vertex operator subalgebra with a different Virasoro element. Then $M$ is a weak $V^{i}$-module. Since $V^i$ is regular, $M$ is a direct sum of simple ordinary $V^{i}$-modules. Note that there are only finitely many simple $V^{i}$-modules up to equivalence. Thus $M$ is a $V^{i}$-module in category $\cal{O}$ of weak $V^i$-modules. Denote the generators of the Virasoro algebra of $V^i$ by $L_i(n).$ Then $L(0)=L_{1}(0)+\cdots +L_{k}(0)$ and $L_{i}(0)$'s commute with each other. This implies that $M$ is in category $\cal{O}$ of weak $V$-modules. So $M$ is completely reducible by Remarks \ref{r3.2} and \ref{r2.5} (iii). $\;\;\;\;\Box$ \begin{prop}\label{p2.6} Let $V$ be a rational vertex operator algebra such that there is a regular vertex operator subalgebra $U$ with the same Virasoro element $\omega$. Then $V$ is regular. \end{prop} {\bf Proof.} Let $M$ be a weak $V$-module. Then $M$ is a weak $U$-module, so that $M$ is a direct sum of simple ordinary $U$-modules. Thus $L(0)$ acts semisimply on $M$. Let $W^{1},\cdots, W^{k}$ be all simple $U$-modules up to equivalence. Then we can write $M=\sum_{i=1}^{k}\oplus_{n\in {\Bbb Z}_{+}}M_{n+h_{i}}$, where $h_{i}$ is the lowest weight of $W^{i}$. Thus $M$ is in the category $\cal{O}$ for $V.$ The complete reducibility of $M$ follows immediately as $V$ is rational.$\;\;\;\;\Box$ \begin{prop}\label{p2.8} Let $V$ be a rational vertex operator algebra such that any nonzero weak $V$-module contains a simple ordinary $V$-submodule. Then $V$ is regular. \end{prop} {\bf Proof.} Let $M$ be any weak $V$-module and let $W$ be the sum of all simple ordinary submodules. We have to prove that $W=M$. If $M\ne W$, $M/W$ is a nonzero weak $V$-module so that by assumption, there is a simple ordinary $V$-submodule $M^{1}/W$ of $M/W$. Then both $W$ and $M^{1}/W$ are in the category $\cal{O}$. Thus $M^{1}$ is in the category $\cal{O}$ and $M^{1}$ is a direct sum of simple ordinary $V$-modules. This contradicts the choice of $W$. $\;\;\;\;\Box$ Now we are ready to show that the vertex operator algebras $L(l,0),$ $L(c_{p,q},0),$ $V^{\natural}$ and $V_L$ are regular. Recall from example (2) that $\frak g$ is a finite-dimensional simple Lie algebra with Cartan subalgebra $\frak h;$ $L(l,0)$ is a vertex operator algebra. We shall denote the corresponding root system by $\Delta.$ \begin{lem}\label{l2.9} There is a basis $\{a^{1},\cdots, a^{m}\}$ for ${\frak g}$ such that for $1\leq i,j\leq m$ we have \begin{eqnarray}\label{eb} [Y(a^{i},z_{1}),Y(a^{i},z_{2})]=0\;\;\mbox{ and }Y(a^{i},z)^{3\ell+1}=0 \end{eqnarray} as operators on $L(\ell,0).$ \end{lem} {\bf Proof.} Let $\alpha\in\Delta$ and $e\in {\frak g}_{\alpha}$. If ${\frak g}$ is of type $A, D,$ or $ E$, it is proved in [MP1] and [DL] that $Y(e,z)^{\ell+1}=0$. In general, it is proved in [L1] and [MP2] that $Y(e,z)^{3\ell+1}=0$. It is well known (cf. [H], [K1]) that there are elements $e_{\alpha}\in {\frak g}_{\alpha}, f_{\alpha}\in {\frak g}_{-\alpha}, h_{\alpha}\in \frak h$ which linearly span a subalgebra naturally isomorphic to $sl_{2}$. Set $\sigma_{\alpha}=e^{(e_{\alpha})_{0}}$ where $(e_{\alpha})_0$ is the component operator of $Y(e_{\alpha},z)$ (cf. (\ref{1/2})) corresponding to $z^{-1}.$ Then $\sigma_{\alpha}$ is an automorphism of the vertex operator algebra $L(\ell,0)$ (see Chap. 11 of [FLM]). A straightforward calculation gives $\sigma_{\alpha}(f_{\alpha})=f_{\alpha}+h_{\alpha}-2e_{\alpha}$. Since $e_{\alpha}, f_{\alpha}, h_{\alpha}$ form a basis of ${\frak g}$ for $\alpha\in \Delta,$ $e_{\alpha}, f_{\alpha}, \sigma_{\alpha}(f_{\alpha})$ also form a basis of ${\frak g}$. It is clear that this basis satisfies condition (\ref{eb}). $\;\;\;\;\Box$ \begin{thm}\label{t2.10} Let $\ell$ be a positive integer. Then the vertex operator algebra $L(\ell,0)$ is regular. \end{thm} {\bf Proof.} By Proposition \ref{p2.8}, it is enough to prove that any nonzero weak $L(\ell,0)$-module $M$ contains a simple $L(\ell,0)$-module. This will be established in three steps. {\bf Claim 1:} {\em There exists a nonzero $u\in M$ such that $(t{\Bbb C}[t]\otimes {\frak g})u=0$.} Set $\frak g(n)=t^n\otimes \frak g$ for $n\ne 0.$ For any nonzero $u\in M$, by the definition of a weak module, ${\frak g}(n)u=0$ for sufficiently large $n.$ So $(t{\Bbb C}[t]\otimes {\frak g})u$ is finite-dimensional. For any $u\in M$, we define $d(u)=\dim (t{\Bbb C}[t]\otimes {\frak g})u$. If there is a $0\ne u\in M$ such that $d(u)=0$, then $(t{\Bbb C}[t]\otimes {\frak g})u=0$. Suppose that $d(u)>0$ for any $0\ne u\in M$. Take $0\ne u\in M$ such that $d(u)$ is minimal. Let $a^{i}$ ($1\le i\le m$) be a basis of ${\frak g}$ satisfying condition (\ref{eb}) and let $k$ be the positive integer such that ${\frak g}(k)u\ne 0$ and ${\frak g}(n)u=0$ whenever $n>k$. By definition of $k$, $a^{i}(k)u\ne 0$ for some $1\le i\le m$. Since $Y(a^{i},z)^{3\ell+1}=0$, by Proposition 13.16 in [DL], $Y_{M}(a^{i},z)^{3\ell+1}=0$. Extracting the coefficient of $z^{-(k+1)(3\ell+1)}$ from $Y_{M}(a^{i},z)^{3\ell+1}u=0$ we obtain $(a^{i}_{k})^{3\ell+1}u=0$. Let $r$ be a nonnegative integer such that $(a^{i}_{k})^{r}u\ne 0$ and $(a^{i}_{k})^{r+1}u=0$. Set $v=(a^{i}_{k})^{r}u$. We will obtain a contradiction by showing that $d(v)<d(u)$. First we prove that if $a_{n}u=0$ for some $a\in {\frak g}, 1\le n\in {\Bbb Z}$, then $a_{n}v=0$. In the following we will show by induction on $m$ that $a_n(a_k^i)^mu=0$ for any $a\in \frak g$ and $m\in \Bbb Z$ nonnegative. If $m=0$ this is immediate by the choice of $u.$ Now assume that the result holds for $m.$ Since $[a,a^{i}]_{k+n}u=0$ (from the definition of $k$) and $a_{n}u=0,$ by the induction assumption that $a_n(a_k^i)^mu=0$ we have: \begin{eqnarray} [a,a^{i}]_{k+n}(a^{i}_{k})^{m}u=0,\;a_{n}(a^{i}_{k})^{m}u=0. \end{eqnarray} Thus \begin{eqnarray} & &a_{n}(a^{i}_{k})^{m+1}u=[a_{n},a^{i}_{k}](a^{i}_{k})^{m}u+a^{i}_{k}a_{n}(a^{i}_{k})^{m}u\nonumber\\ & &\ \ \ \ =[a,a^{i}]_{k+n}(a^{i}_{k})^{m}u+a^{i}_{k}a_{n}(a^{i}_{k})^{m}u\nonumber\\ & &\ \ \ \ =0, \end{eqnarray} as required. In particular, we see that $a_{n}v=a_{n}(a^{i}_{k})^{r}u=0$. Therefore, $d(v)\le d(u)$. Since $a^{i}_{k}v=0$ and $a^{i}_{k}u\ne 0$, we have $d(v)< d(u)$. {\bf Claim 2:} {\em There is a nonzero $u\in M$ such that ${\frak g}(n)u=0$ for $n>0$ and ${\frak g}_+u=0$ where $\frak g_+=\sum_{\alpha\in \Delta_+}\frak g_{\alpha}$ for a fixed positive root system $\Delta_+.$} Set \begin{eqnarray} \Omega(M)=\{u\in M| (t{\Bbb C}[t]\otimes {\frak g})u=0\}. \end{eqnarray} Then $\Omega(M)$ is a nonzero ${\frak g}$-submodule of $M$ by Claim 1. Let $0\ne e_{\theta}\in {\frak g}_{\theta}$ where $\theta$ is the longest positive root in $\Delta$. Then $Y_{M}(e_{\theta},z)^{\ell+1}=0$ (see [DL] and [FZ]). Extracting the coefficient of $z^{-\ell-1}$ from $Y_{M}(e_{\theta},z)^{\ell+1}\Omega(M)=0$, we obtain $e_{\theta}^{\ell+1}\Omega(M)=0$. By Proposition 5.1.2 of [L1] $\Omega(M)$ is a direct sum of finite-dimensional irreducible ${\frak g}$-modules. Then any highest weight vector for ${\frak g}$ in $\Omega(M)$ meets our need. {\bf Claim 3:} {\em Any lowest weight vector for $\hat{{\frak g}}$ in $M$ generates a simple $L(\ell,0)$-module.} Let $u$ be a lowest weight vector for $\hat{{\frak g}}$ in $M$. Extracting the constant term from $Y_{M}(e_{\theta},z)^{\ell+1}u=0$, we obtain $(e_{\theta})_{-1}^{\ell+1}u=0$. Then $u$ generates an integrable highest weight $\hat{{\frak g}}$-module. It follows from [K1] that $u$ generates an irreducible $\hat{{\frak g}}$-module of level $\ell$. Since any submodule of $M$ for the affine Lie algebra is a submodule of $M$ for $L(\ell,0)$, such $u$ generates a simple $L(\ell,0)$-module.$\;\;\;\;\Box$ \begin{rem}\label{r2.3'} This theorem has been proved in [DLM2] under the assumption that $t\Bbb C[t]\otimes \frak h$ acts locally nilpotently on any weak module. See Proposition 5.6 in [DLM2]. \end{rem} \begin{rem} Recall that a $\hat{{\frak g}}$-module $M$ is called {\em restricted } (cf. [K1]) if for any $u\in M$, there is an integer $k$ such that $(t^{n}\otimes {\frak g})u=0$ for $n>k;$ $M$ is called an {\em integrable} module if the Chevalley generators $e_i,f_i$ of $\hat{{\frak g}}$ act locally finitely on $M$ [K2] (note that in the definition of integrable module, we do not assume that the action of $\frak h$ is semisimple). At affine Lie algebra level, Proposition \ref{p2.8} and Theorem \ref{t2.10} essentially assert that any restricted integrable $\hat{{\frak g}}$-module is a direct sum of irreducible highest weight integrable $\hat{{\frak g}}$-modules. \end{rem} Next we turn our attention to the vertex operator algebras $L(c_{p,q},0).$ First we recall some results from [DL]. Let $V$ be a vertex operator algebra and $M$ be a weak $V$-module. Then for any $u,v\in V$ we have \begin{eqnarray}\label{ef} Y(u_{-1}v,z)=Y(u,z)^{-}Y(v,z)+Y(v,z)Y(u,z)^{+}, \end{eqnarray} where \begin{eqnarray} Y(u,z)^{+}=\sum_{n\ge 0}u_{n}z^{-n-1}, \; Y(u,z)^{-}=\sum_{n<0}u_{n}z^{-n-1}. \end{eqnarray} Note that $Y(u,z)=Y(u,z)^{+}+Y(u,z)^{-}$ and that $Y(u,z)^{+}$ (reps. $Y(u,z)^{-}$) involves only nonpositive (resp. nonnegative) powers of $z$. For convenience we will write $c=c_{p,q}.$ For any nonnegative integer $n$, we set $\omega^{(n)}={1\over n!}L(-1)^{n}\omega$. Then $L(-n-2)=\omega^{(n)}_{-1}.$ We need the following lemmas. \begin{lem}\label{l2.12} Let $M$ be a weak $L(c,0)$-module and $u\in M.$ Let $k$ be a positive integer such that $L(k)u\ne 0$ and that $L(n)u=0$ whenever $n>k$. Then for any nonnegative integers $n_{1},..., n_{r}$ the lowest power of $z$ in $Y_{M}(\omega^{(n_{1})}_{-1}\cdots \omega^{(n_{r})}_{-1}{\bf 1},z)u$ (in the sense that the coefficients of $z^m$ is zero whenever $m$ is smaller than the lowest power) is $-r(k+2)-n_{1}-\cdots -n_{r}$ with coefficient $\displaystyle{\prod_{i=1}^{r}{-k-2\choose n_{i}}L(k)^{r}u}.$ \end{lem} {\bf Proof.} We prove this lemma by induction on $r$. If $r=1$, we have: \begin{eqnarray} & &Y_M(\omega^{(n_{1})}_{-1}{\bf 1},z)u=Y_M(\omega^{(n_{1})},z)u\nonumber\\ & &\ \ \ \ \ =\frac{1}{n_1!}\left({d\over dz}\right)^{n_{1}}Y_M(\omega,z)u\nonumber\\ & &\ \ \ \ \ =\sum_{n\in {\Bbb Z}}{-n-2\choose n_{1}}z^{-n-2-n_{1}}L(n)u \nonumber\\ & &\ \ \ \ \ =\sum_{n\le k}{-n-2\choose n_{1}}z^{-n-2-n_{1}}L(n)u. \end{eqnarray} Then the lowest power of $z$ is $-(k+2)-n_{1}$ with a coefficient ${-k-2\choose n_{1}}L(k)u$. That is, the lemma holds for $r=1$. Suppose that this lemma holds for some positive integer $r$. By formula (\ref{ef}) we have: \begin{eqnarray} & &Y_{M}(\omega^{(n_{1})}_{-1}\cdots \omega^{(n_{r})}_{-1}\omega^{(n_{r+1})}_{-1}{\bf 1},z)u=Y_{M}(\omega^{(n_{1})},z)^{-}Y_{M}(\omega^{(n_{2})}_{-1}\cdots \omega^{(n_{r+1})}_{-1}{\bf 1},z) u\nonumber\\ & &\ \ \ \ \ \ +Y_{M}(\omega^{(n_{2})}_{-1}\cdots \omega^{(n_{r+1})}_{-1}{\bf 1},z) Y_{M}(\omega^{(n_{1})},z)^{+}u.\label{a2.15} \end{eqnarray} Since $Y_{M}(\omega^{(n_{1})},z)^{-}$ involves only nonnegative powers of $z$, it follows from the inductive assumption, the lowest power of $z$ in the first term of the right hand side of (\ref{a2.15}) is $-r(k+2)-n_{2}-\cdots -n_{r+1}$. It is easy to observe that for any $v$ in the algebra, $$Y_M(L(-1)v,z)^+=\frac{d}{dz}Y_M(v,z)^+.$$ Thus \begin{eqnarray*} & &\ \ \ \ Y_{M}(\omega^{(n_{2})}_{-1}\cdots \omega^{(n_{r+1})}_{-1}{\bf 1},z) Y_{M}(\omega^{(n_{1})},z)^{+}u\\ & &=\frac{1}{n_1!}Y_{M}(\omega^{(n_{2})}_{-1}\cdots \omega^{(n_{r+1})}_{-1} {\bf 1},z)\left(\frac{d}{dz}\right)^{n_1}Y_{M}(\omega,z)^{+}u\\ & &=\sum_{n=-1}^k{-n-2\choose n_1}Y_{M}(\omega^{(n_{2})}_{-1}\cdots \omega^{(n_{r+1})}_{-1}{\bf 1},z)L(n)uz^{-n-2-n_1}\\ & &=\sum_{n=0}^k{-n-2\choose n_1}Y_{M}(\omega^{(n_{2})}_{-1}\cdots \omega^{(n_{r+1})}_{-1}{\bf 1},z)L(n)uz^{-n-2-n_1}\\ & &\ \ \ \ \ +{-1\choose n_1}L(-1)Y_{M}(\omega^{(n_{2})}_{-1}\cdots \omega^{(n_{r+1})}_{-1}{\bf 1},z)uz^{-1-n_1}\\ & &\ \ \ \ \ +{-1\choose n_1}\left(\frac{d}{dz}Y_{M}(\omega^{(n_{2})}_{-1}\cdots \omega^{(n_{r+1})}_{-1}{\bf 1},z)\right)uz^{-1-n_1}. \end{eqnarray*} Note that $L(m)L(n)u=(m-n)L(m+n)u+L(n)L(m)u=0$ for $0\leq n\leq k$ and $m>k.$ Applying the inductive hypothesis to $L(n)u$ we see that the lowest power of $z$ in $$\sum_{n=0}^k{-n-2\choose n_1}Y_{M}(\omega^{(n_{2})}_{-1}\cdots \omega^{(n_{r+1})}_{-1}{\bf 1},z)L(n)uz^{-n-2-n_1}$$ is $-(r+1)(k+2)-n_1-\cdots -n_{r+1}$ with coefficient $\displaystyle{\prod_{i=1}^{r+1}{-k-2\choose n_i}L(k)^{r+1}u.}$ Also by the induction assumption the lowest power of $z$ in $L(-1)Y_{M}(\omega^{(n_{2})}_{-1}\cdots \omega^{(n_{r+1})}_{-1}{\bf 1},z)uz^{-1-n_1}$ is $-r(k+2)-n_1-\cdots -n_{r+1}-1$ and the lowest power of $z$ in $\left(\frac{d}{dz}Y_{M}(\omega^{(n_{2})}_{-1}\cdots \omega^{(n_{r+1})}_{-1}{\bf 1},z)\right)uz^{-1-n_1}$ is $-r(k+2)-n_1-\cdots -n_{r+1}-2.$ Thus the lowest power of $z$ in the second term of right hand side of (\ref{a2.15}) is $-(r+1)(k+2)-n_{1}-\cdots -n_{r+1}$ with coefficient $\displaystyle{\prod_{i=1}^{r+1}\left(\begin{array}{c}-k-2\\n_{i}\end{array}\right) L(k)^{r+1}u},$ as desired. $\;\;\;\;\Box$ \bigskip Let $V$ be a vertex operator algebra and let $A(V)$ (which is a certain quotient space of $V$ modulo a subspace $O(V)$) be the corresponding associative algebra defined in [Z]. We refer the reader to [Z] for details. Recall from [DLM1] or [L2] that for any weak $V$-module $M,$ $\Omega(M)$ consists of vectors $u\in M$ such that $a_{m}u=0$ for any homogeneous element $a\in V$ and for any $m>{\rm wt}\,a-1$. In other words, $u\in \Omega(M)$ if and only if $z^{m}Y_{M}(a,z)u\in M[[z]]$ for any homogeneous element $a\in V$ and for any $m>{\rm wt}a-1$. The following result can be found in [DLM1], [L2] and [Z]. \begin{lem}\label{al} (1) $\omega+O(V)$ is in the center of $A(V).$ (2) $A(V)$ is semisimple if $V$ is rational. (3) $\Omega(M)$ is an $A(V)$-module under the action $a+O(V)\mapsto a_{{\rm wt}\,a-1}$ for homogeneous $a\in V.$ \end{lem} Now we take $V=L(c,0).$ Set $\bar{\Omega}(M)=\{ u\in M|L(n)u=0\;\mbox{ for any }n>0\}$. Then it is clear that $\Omega(M)\subseteq \bar{\Omega}(M)$. \begin{lem}\label{l2.13} Let $M$ be a weak $L(c,0)$-module. Then $\Omega(M)=\bar{\Omega}(M)$. \end{lem} {\bf Proof.} It suffices to prove that $a_{m}u=0$ for any $u\in \bar{\Omega}(M)$ and for any homogeneous element $a\in L(c,0)$ whenever $m>{\rm wt}a-1$. We shall prove this by induction on the weight of $a$. If ${\rm wt}a=0$, $a={\bf 1}$. Since ${\bf 1}_{m}=0$ for $m\ge 0$, there is nothing to prove. Suppose that $a_{m}u=0$ for any homogeneous element $a\in L(c,0)$ of weight less than $n$ and for any $m>{\rm wt}a-1$. Let $b\in L(c,0)$ be a homogeneous element of weight $n$ and let $m\in {\Bbb Z}$ such that $m>{\rm wt}b-1$. Let $a\in L(c,0)$ be any homogeneous element of weight less than $n$, let $k$ be any positive integer and let $m>{{\rm wt}}\,(L(-k)a)-1$ $(={\rm wt}\,a+k-1)$. Then from the Jacobi identity (\ref{jacobi}) we have: \begin{eqnarray} & &(L(-k)a)_{m}u={\rm Res}_{z_{0}}{\rm Res}_{z_{2}}z_{0}^{1-k}z_{2}^{m}Y_{M}(Y(\omega,z_{0})a,z_{2})u\nonumber\\ & &\ \ ={\rm Res}_{z_{1}}{\rm Res}_{z_{0}}{\rm Res}_{z_{2}}z_{0}^{1-k}z_{2}^{m}\cdot \nonumber\\ & &\ \ \ \cdot\left( z_{0}^{-1}\delta\left(\frac{z_{1}-z_{2}}{z_{0}}\right)Y_{M}(\omega,z_{1})Y_{M}(a,z_{2})u -z_{0}^{-1}\delta\left(\frac{z_{2}-z_{1}}{-z_{0}}\right)Y_{M}(a,z_{2})Y_{M}(\omega,z_{1})u\right) \nonumber\\ & &\ \ ={\rm Res}_{z_{1}}{\rm Res}_{z_{2}}(z_{1}-z_{2})^{1-k}z_{2}^{m}Y_{M}(\omega,z_{1})Y_{M}(a,z_{2})u \nonumber\\ & &\ \ \ -{\rm Res}_{z_{1}}{\rm Res}_{z_{2}}(-z_{2}+z_{1})^{1-k}z_{2}^{m}Y_{M}(a,z_{2})Y_{M}(\omega,z_{1})u. \end{eqnarray} Since $m>{\rm wt}\,(L(-k)a)-1={\rm wt}a+k-1>{\rm wt}a-1$, we have: $${\rm Res}_{z_{1}}{\rm Res}_{z_{2}}(z_{1}-z_{2})^{1-k}z_{2}^{m}Y_{M}(\omega,z_{1})Y_{M}(a,z_{2})u=0.$$ For the second term, we have: \begin{eqnarray} & &\ \ \ -{\rm Res}_{z_{1}}{\rm Res}_{z_{2}}(-z_{2}+z_{1})^{1-k}z_{2}^{m}Y_{M}(a,z_{2})Y_{M}(\omega,z_{1})u \nonumber\\ & &=-{\rm Res}_{z_{2}}(-1)^{1-k}z_{2}^{m+1-k}Y_{M}(a,z_{2})L(-1)u -{\rm Res}_{z_{2}}(-1)^{-k}(1-k)z_{2}^{m-k}Y_{M}(a,z_{2})L(0)u\nonumber\\ & &=-{\rm Res}_{z_{2}}(-1)^{1-k}z_{2}^{m+1-k}L(-1)Y_{M}(a,z_{2})u +{\rm Res}_{z_{2}}(-1)^{1-k}z_{2}^{m+1-k}{d\over dz_{2}}Y_{M}(a,z_{2})u\nonumber\\ & &\ \ \ -{\rm Res}_{z_{2}}(-1)^{-k}(1-k)z_{2}^{m-k}Y_{M}(a,z_{2})L(0)u\nonumber\\ & &=-{\rm Res}_{z_{2}}(-1)^{1-k}z_{2}^{m+1-k}L(-1)Y_{M}(a,z_{2})u -{\rm Res}_{z_{2}}(-1)^{1-k}(m+1-k)z_{2}^{m-k}Y_{M}(a,z_{2})u\nonumber\\ & &\ \ \ -{\rm Res}_{z_{2}}(-1)^{-k}(1-k)z_{2}^{m-k}Y_{M}(a,z_{2})L(0)u\nonumber\\ & &=(-1)^kL(-1)a_{m+1-k}u+(-1)(m+1-k)a_{m-k}u+(-1)^{k-1}a_{m-k}L(0)u.\label{2.17} \end{eqnarray} Since $L(0)u\in \bar{\Omega}(M)$ and $m-k>{\rm wt}a-1$ all the three terms in (\ref{2.17}) are zero by the inductive hypothesis. Thus $(L(-k)a)_mu=0.$ Note that $b$ is a linear combination of all $L(-k)a$, where ${\rm wt}a<n$ and $k$ is a positive integer. This shows $b_{m}\bar{\Omega}(M)=0$ for $m>{\rm wt}\,b-1,$ as desired. $\;\;\;\;\Box$ Now we are in a position to prove \begin{thm}\label{t2.11} The vertex operator algebra $L(c,0)$ associated with the lowest weight irreducible module for the Virasoro algebra with central charge $c=c_{p,q}$ is regular. \end{thm} {\bf Proof:} By Proposition \ref{p2.8}, it is enough to prove that any nonzero weak $L(c,0)$-module $M$ contains a simple $L(c,0)$-module. {\bf Claim 1:} {\em The space $\Omega(M)$ is not zero.} For any $0\ne u\in M$, we define $l(u)$ to be the integer $k$ such that $L(k)u\ne 0$ and $L(n)u=0$ whenever $n>k$. Since $L(n)u\ne 0$ for some $n$ (because $c\ne 0$), $l(u)$ is well-defined. Suppose that $\Omega(M)=0$. Then by Lemma \ref{l2.13} $l(u)\ge 1$ for any $0\ne u\in M$. Let $0\ne u\in M$ such that $l(u)=k$ is minimal. It is well known [FF] that there are two singular vectors in the Verma module $M(c,0)$ for the Virasoro algebra. One singular vector is $L(-1){\bf 1}$ and the other is: \begin{eqnarray} v=L(-2)^{pq}{\bf 1}+\sum a_{n_{1},\cdots, n_{r}}\omega^{(n_{1})}_{-1}\cdots \omega^{(n_{r})}_{-1}{\bf 1}, \end{eqnarray} where the sum is over some $(n_{1},\cdots, n_{r})\in {\Bbb Z}_{+}^{r}$ such that $2pq=2r+n_{1}+\cdots +n_{r}$ and $n_{1}+\cdots +n_{r}\ne 0$. By Lemma \ref{l2.12}, the lowest power of $z$ in $$Y_{M}(L(-2)^{pq}{\bf 1},z)u=Y_{M}((\omega_{-1})^{m}{\bf 1},z)u$$ is $-pq(k+2)$ with $L(k)^{pq}u$ as its coefficient and the lowest power of $z$ in $$Y_{M}(\omega^{(n_{1})}_{-1}\cdots \omega^{(n_{r})}_{-1}{\bf 1},z)u$$ is greater than $-pq(k+2)$ for any nonnegative integers $n_{1},\cdots, n_{r}$ such that $2pq=2r+n_{1}+\cdots +n_{r}$ and $n_{1}+\cdots +n_{r}\ne 0$. Thus the coefficient of $z^{-pq(k+2)}$ in $Y_{M}(v,z)u$ is $L(k)^{pq}u$. Since $v=0$ in $L(c,0)$ we have $Y_{M}(v,z)=0$. In particular the coefficient $L(k)^{pq}u$ of $z^{-pq(k+2)}$ in $Y_M(v,z)$ is zero. Let $s$ be the nonnegative integer such that $L(k)^{s}u\ne 0$ and $L(k)^{s+1}u=0$ and set $u'=L(k)^{s}u$. Then it is clear that $l(u')<l(u)$. This is a contradiction. {\bf Claim 2:} {\em Any weak $L(c,0)$-module $M$ contains a simple ordinary $L(c,0)$-module.} Since $L(c,0)$ is rational, Lemma \ref{al} tells us that $A(L(c,0))$ is semisimple and that the central element $\omega+O(L(c,0))$ acts semisimply on $\Omega(M)$ as $L(0).$ Since $\Omega(M)$ is nonzero by Claim 1 we can take $0\ne u\in \Omega(M)$ such that $L(0)u=hu$ where $h\in \Bbb C.$ Again since $L(c,0)$ is rational [W], $u$ generates a simple (ordinary) $L(c,0)$-module. The proof is complete. $\;\;\;\;\Box$ \begin{coro}\label{cm} The moonshine module vertex operator algebra $V^{\natural}$ is regular. \end{coro} {\bf Proof.} From [DMZ], $V^{\natural}$ contains $L({1\over 2},0)^{\otimes 48}$ as a vertex operator subalgebra. Then the result follows from Theorem \ref{t2.11}, Propositions \ref{p2.5} and \ref{p2.6}. $\;\;\;\;\Box$ Finally we discuss the complete reducibility of weak $V_L$-modules for an even lattice $L.$ We refer the reader to [FLM] and [D1] for the construction of $V_L$ and related notations. Let $M$ be any weak $V_L$-module. Define the vacuum space $$\Omega_M=\{u\in M| \alpha(i)u=0\ {\rm for}\ \alpha \in L,\ i>0\}.$$ \begin{lem}\label{p2.14} Let $L$ be an even lattice. Then for any weak $V_{L}$-module $M,$ $\Omega_M\ne 0.$ \end{lem} {\bf Proof.} For $u\in M$ then $A_u={\rm span}\{\alpha (n)u|\alpha\in L, n>0\}$ is finite-dimensional as $\alpha (n)u=0$ if $n$ is sufficiently large and as the rank of $L$ is finite. Set $d(u)=\dim A_u.$ Note that $d(u)=0$ if and only if $u\in \Omega_M.$ So it is enough to show that $d(u)=0$ for some nonzero $u\in M.$ Assume this is false, and take $0\ne u\in M$ such that $d(u)$ is minimal. Let $k$ be the smallest positive integer such that $\alpha (k)u\ne 0$ and $\beta (n)u=0$ whenever $n>k$ for some $\alpha \in L$ and all $\beta \in L.$ Let $a\in \hat L$ such that $\bar a=\alpha .$ Then from the formula (3.4) of [D1] we have $$\frac{d}{dz}Y(\iota(a),z)=Y(L(-1)\iota(a),z)=Y(\alpha(-1)\iota(a),z) =\alpha(z)^-Y(\iota(a),z)+Y(\iota(a),z)\alpha(z)^+$$ where $$\alpha(z)^-=\sum_{n<0}\alpha(n)z^{-n-1},\ \ \alpha(z)^+=\sum_{n\geq 0}\alpha(n)z^{-n-1}.$$ Clearly the submodule generated by $u$ is not zero. Note that the vertex algebra $V_L$ is simple (see [D1]). By Proposition 11.9 of [DL], $Y(\iota(a),z)u\ne 0.$ Let $r$ be an integer such that $\iota(a)_{r+m}u=0$ and $\iota(a)_{r}u\ne 0$ for any positive integer $m.$ Thus the lowest power of $z$ in $$\frac{d}{dz}Y(\iota(a),z)u=-\sum_{m\leq r}(m+1)\iota(a)_muz^{-m-2}$$ is at most $-r-2.$ It is obvious that the lowest power of $z$ in $ \alpha(z)^-Y(\iota(a),z)u$ is at most $-r-1.$ Use the following commutator formula which is a result from the Jacobi identity \begin{equation}\label{gr} [\beta(m),\iota(a)_n]=\<\alpha ,\beta \>\iota(a)_{m+n} \end{equation} to obtain $$\iota(a)_m\alpha(n)u=-\<\alpha ,\alpha \>\iota(a)_{m+n}u+\alpha(n)\iota(a)_mu=0$$ if $m>r$ and $n\geq 0.$ This gives $$Y(\iota(a),z)\alpha(z)^+u= \sum_{m\leq r}\sum_{n=0}^k\iota(a)_m\alpha (n)z^{-m-n-2}.$$ Thus the coefficient $\iota(a)_{r}\alpha (k)u$ of $z^{-r-k-2}$ in the formula above is zero as $k$ is positive. This shows by (\ref{gr}) again that $\alpha(n)\iota(a)_ru=0$ for any positive integer greater than or equal to $k.$ Note from (\ref{gr}) that if $\beta (m)u=0$ for positive $m$ then $\beta (m)\iota(a)_ru=0.$ Thus $d(\iota(a)_ru)< d(u).$ This is a contradiction. $\;\;\;\;\Box$ \begin{thm}\label{tvl} Let $L$ be an even lattice. Then any weak $V_L$-module is completely reducible and any simple weak $V_L$-module is isomorphic to $V_{L+\beta}$ for some $\beta $ in the dual lattice of $L.$ In particular, $V_L$ is regular if $L$ is positive definite. \end{thm} {\bf Proof.} By Lemma \ref{p2.14}, $\Omega_M\ne 0$ for a weak $V_L$-module $M$. It is proved in [D1] that if $M$ is also simple then it is necessarily isomorphic to $V_{L+\beta}$ for some $\beta $ in the dual lattice of $L.$ So it remains to show the complete reducibility of any weak $V_L$-module $M.$ Let $W$ be the sum of all simple submodules of $M.$ Assume that $M'=M/W$ is not zero. Then $\Omega_{M'}\ne 0.$ It is essentially proved in [D1] that $M'$ contains a simple module $W^1/W$ (here $W^1$ is a weak $V_L$-submodule of $M$ which contains $W$) generated by $w^1+W$ where $w^1$ is a common eigenvector for the operators $\alpha (0)$ for $\alpha \in L.$ It follows from the proof of Theorem 3.1 of [D1] that the submodule of $M$ generated by $w^1$ is simple. Thus $W^1$ is a sum of certain simple submodules of $V_L.$ This is a contradiction. $\;\;\;\;\Box$ At this point we have proved that almost all known rational vertex operator algebras are regular. We conclude this paper by presenting the following conjecture: \begin{conj} Any rational vertex operator algebra is regular. \end{conj}
\section{Introduction} As was recently noted by Blaylock, Seiden, and Nir\cite{Blaylock} due to final state interaction (FSI) a term proportional to $\Delta M ~t~ e^{-\Gamma t}$ may appear in the rate of wrong sign $D$ decays even in the absence of CP violation. Moreover, in some extensions of Standard Model which have large values of both $\Delta M$ and significant CP violation, a similar term may arise. Blaylock {\it et al.} have suggested that a value of $\Delta M$ larger than the present experimental limit can be accomodated if one of these previously neglected terms destructively interferes with the other time dependent terms which arise from mixing (proportional to $t^2 ~e^{-\Gamma t}$) and from doubly Cabibbo suppressed decays (DCSD) (proportional to $e^{-\Gamma t}$). They suggest that this may invalidate the use of existing limits from time dependent mixing studies at fixed target experiments \cite{E691} , \cite{E791} to constrain extensions of the Standard Model. Below, we give expressions for the time dependence in the general case and then attempt to estimate the maximum size of the terms proportional to $t e^{-\Gamma t}$. \section{Formalism for Mixing} We follow the notation of references \cite{Blaylock},\cite{liu1},\cite{liu2}. Let the mass eigenstates be $D_S$, $D_L$. Then \[ |D_S> = p |D^0> + q |\bar{D}^0> \] \[ |D_L> = p |D^0> - q |\bar{D}^0> \] In the limit of no CP violation, $p$=$q$=$1/\sqrt{2}$. Let $\Delta M= M_L- M_S$ and $\Delta \Gamma = \Gamma_L-\Gamma_S$ denote the mass difference and lifetime difference, respectively. Let A denote the amplitude for $<f|H|D^0>$, B the amplitude for $<f|H|\bar{D}^0>$. Let $\displaystyle\lambda={p \over q}{A\over B}$ and $\displaystyle\bar{\lambda}={q \over p}{\bar{A}\over \bar{B}}$. The decay rate is then given by \begin{equation} \Gamma(D^0(t)\to K^+\pi^-) = {e^{-\Gamma t} \over 4} |B|^2 |{q\over p}|^2 \{ 4 |\lambda|^2 + ({\Delta M}^2 + {{\Delta \Gamma}^2 \over 4}) t^2 + 2 Re(\lambda) \Delta\Gamma t + 4 Im(\lambda) \Delta M t \} \label{mixeqn1} \end{equation} up to terms of order $t^2$~\cite{Blaylock}. The decay rate for the charge conjugate reaction is given by the same expression replacing $\lambda$ with $\bar{\lambda}$, $B$ with $\bar{B}$, and $q/p$ by $p/q$. \begin{equation} \Gamma(\bar{D}^0(t)\to K^-\pi^+) = {e^{-\Gamma t} \over 4} |\bar{B}|^2 |{p\over q}|^2 \{ 4 |\bar{\lambda}|^2 + ({\Delta M}^2 + {{\Delta \Gamma}^2 \over 4}) t^2 + 2 Re(\bar{\lambda}) \Delta\Gamma t + 4 Im(\bar{\lambda}) \Delta M t \} \label{mixeqn2} \end{equation} In the past, it was assumed that the term proportional to $\Delta M ~t$ changes sign when averaging over a sample with equal numbers of $D^0$ and $\bar{D^0}$ mesons\cite{Bigi},\cite{Browder}. This assumption is not correct in general as was noted in Reference\cite{Blaylock}. The previous experimental analyses\cite{E691}, \cite{Browder}, \cite{E791} considered the deterioration of the limit in the case when the term proportional to $\Delta \Gamma ~t$ interfered destructively with the mixing and DCSD components. The Standard Model expectation for $\Delta \Gamma$ is many orders of magnitude below the current experimental sensitivity so this interference scenario is very unlikely. In most new physics scenarios which would give $r_{mix}\sim O(10^{-3})$, $\Delta\Gamma$ is not enhanced whereas values of $\Delta M$ much larger than those expected from the Standard Model are possible. It is also possible to experimentally verify that $\Delta \Gamma$ can be neglected by measuring the $D$ meson lifetime in a CP eigenstate e.g. $D^0\to K^-K^+$ and comparing to the lifetime in $D^0\to K^-\pi^+$\cite{liu1},\cite{liu2}. We now consider equations ~(\ref{mixeqn1}),~(\ref{mixeqn2}) in the following situation. Let \[ {p\over q}=\beta e^{i \phi} \] \[ {A\over B}=\alpha e^{i \delta} \] and $\displaystyle \alpha^2= {\Gamma(D^0\to K^+\pi^-)\over \Gamma(D^0\to K^-\pi^+)}$. The phase $\phi$ is due to CP violation in the mass matrix. A non-zero value of $\delta$ may arise if the amplitudes A and B have different FSI. Alternatively, if there are complex contributions to one of the amplitudes (e.g. A) that are not present in the other (e.g. B), this can lead to an overall phase in $\displaystyle{A\over B}$. We have assumed that there is no direct CP violation in the amplitudes (and hence e.g. $\Gamma(D^0\to K^- \pi^+)=\Gamma(\bar{D}^0\to K^+ \pi^-)$)\cite{phase}. In addition, we neglect the small phase in $A/B$ from the CKM matrix, which is approximately $A^2\lambda^4\eta$ in the Wolfenstein parameterization and which lies in the range $(2.3-5.3)\times 10^{-4}$. The decay rate for wrong sign $D^0$ decays to $K^+\pi^-$ is given by \begin{eqnarray*} \Gamma(D^0(t)\to K^+\pi^-) & = & {e^{-\Gamma t} \over 4} |B|^2 \times \\ & & {1\over \beta^2} \{ 4 \alpha^2 \beta^2 + ({\Delta M}^2 + {{\Delta \Gamma}^2 \over 4}) t^2 \\ & & + 2 \alpha \beta \cos(\phi+\delta)(\Delta \Gamma t) +4 \sin(\phi+\delta)\alpha\beta \Delta M t \} \end{eqnarray*} The corresponding rate for the charge conjugate reaction is obtained by replacing $\phi$ the phase from CP violation with $-\phi$ and by changing $\beta$ to $1/\beta$ \begin{eqnarray*} \Gamma(\bar{D}^0 (t)\to K^-\pi^+) &= & {e^{-\Gamma t} \over 4} |\bar{B}|^2 \times \\ & & {\beta^2} \{ 4 {\alpha^2 \over \beta^2} + ({\Delta M}^2 + {{\Delta \Gamma}^2 \over 4}) t^2 \\ & & + 2 {\alpha \over \beta} \cos(-\phi+\delta)(\Delta \Gamma t) +4 \sin(-\phi+\delta){\alpha\over\beta} \Delta M t \} \end{eqnarray*} In the experimental analyses, the time dependent rate integrated over both types of particles is used: $$ \Gamma (D^0(t)\to K^+\pi^-) + \Gamma (\bar{D}^0(t)\to K^-\pi^+) $$ This rate, which will be denoted by $\Gamma(D^0(t)+\bar{D}^0(t))$, is given by \begin{eqnarray} \Gamma(D^0(t)+\bar{D}^0(t)) &= & 2 F(t) \times \\ & & \{ 4 \alpha^2 + {1\over 2}(\beta^2+{1\over \beta^2}) ({\Delta M}^2 + {{\Delta \Gamma}^2 \over 4}) t^2 \nonumber \\ & & + \alpha (\beta\cos(-\phi+\delta) +{1\over\beta}\cos(\phi+\delta))\Delta \Gamma ~t \nonumber \\ & & + 2 \alpha (\beta\sin(-\phi+\delta) + {1\over\beta}\sin(\phi+\delta))\Delta M ~t \} \nonumber \end{eqnarray} where $\displaystyle F(t)={1\over 4} e^{-\Gamma t} ~|B|^2$. \bigskip \section{Effects of FSI and CP Violation} Two scenarios are considered in what follows. First the case of no CP violation but significant final state interactions (FSI) and then the case of both large CP violation and significant final state interaction are examined. \subsection{Effects of FSI} In the first scenario, consider the case of large mixing with $\Delta M>>\Delta M_{SM}$, the value in the Standard Model. Assume that this does not lead to an enhancement of $\Delta \Gamma$ i.e. $\Delta\Gamma_{SM}=\Delta\Gamma<<\Delta M$ and allow for non-zero $\delta$ but no CP violation ($\phi=0, \beta=1$). The above equation then reduces to \begin{equation} \Gamma(D^0(t)+\bar{D}^0(t)) = 2 F(t) \{ 4 \alpha^2 + ({\Delta M}^2 + {{\Delta \Gamma}^2 \over 4}) t^2 +4 \alpha (\sin(\delta))\Delta M ~t \} \end{equation} In order to determine the size of the new term proportional to $\Delta M ~t$, the values of the phase difference $\delta$ are considered in various models. This will allow an estimate of the additional experimental systematic error that is incurred from ignoring FSI. This phase difference $\delta$ is zero in the limit of exact SU(3) symmetry. The values of $\delta$ from various models are given in Table~\ref{Tbdelta}. Large values of the phase $\delta$ occur when SU(3) breaking is largest. We use the experimental result from CLEO~II for $D^0\to K^+ \pi^-$ ($\alpha^2=0.0077\pm 0.0025 \pm 0.0025$) and assume that it is entirely due to DCSD. This is found numerically to give the most conservative upper limit on the size of the interference effect. In general, the amplitudes for the $D^0\to K^- \pi^+$ and $D^0\to K^+ \pi^-$ can be written as: \begin{equation} A(D^0\to K^-\pi^+) = e^{i \delta_3} [(A_1 + C) e^{i (\delta_1-\delta_3)} + A_3 ] \end{equation} \begin{equation} A(D^0\to K^+\pi^-) = -\theta_c^2 e^{i \delta_3} [ (\tilde{A}_1 + \tilde{C}) e^{i (\delta_1-\delta_3)} + \tilde{A}_3 ] \end{equation} where $A_1$ and $A_3$ are the quark decay contributions into $I=1/2$ and $I=3/2$ final states respectively. C is the W-exchange contribution and $\delta_1$ and $\delta_3$ are the FSI phases. $\tilde{A_i}$, $\tilde{C}$ are the corresponding DCSD amplitudes after the CKM factor $-\theta_c^2$ has been factored out. The phase shifts in a given isospin eigenstate for particles and antiparticles are identical by CPT invariance (which we assume as stated explicitly). The phase $\delta$ vanishes if two conditions are satisfied: (i) $\delta_1-\delta_3= \tilde{\delta}_1-\tilde{\delta}_3$ and (ii) $A_3/A_1=\tilde{A_3}/\tilde{A_1}$. The first condition follows from CPT invariance and the second is satisfied if $SU(3)$ symmetry holds. Hence if $SU(3)$ is an approximate symmetry, the phase $\delta$ should be small. The models used have been tuned to reproduce the observed magnitude of $SU(3)$ breaking in $D$ decays. To obtain more information, we turn to the detailed model fits. \subsection{Details of the Models} In the model of Chau and Cheng, \begin{equation} A_1 \cong 0.82, ~A_3\cong 0.16,~C\cong -0.13 \end{equation} \begin{equation} \tilde{A}_1 \cong 1.14, ~\tilde{A}_3\cong 0.33, ~\tilde{C}\approx C \end{equation} and \begin{equation} \delta_1-\delta_3 \approx 90^{0},~ \delta_3\approx 0. \end{equation} Then \begin{equation} A(D^0\to K^-\pi^+)\cong (0.72) e^{i ~76^{0}} \end{equation} \begin{equation} A(D^0\to K^+\pi^-)\cong -\theta_c^2 (1.01) e^{i ~72^{0}} \end{equation} This yields a phase difference between the two decay modes of $\delta=4^{0}$. If the W-exchange contribution $C$ is omitted, the phase difference becomes $\delta=5^{0}$. In the model of Buccella {\it et al.}, one has \begin{equation} {A}_1 \cong 4.35,~ {A}_3\cong -2.3, ~{C}\approx -0.5 \end{equation} \begin{equation} \tilde{A}_1 \cong 5.2, ~\tilde{A}_3\cong -2.3, ~\tilde{C}\approx -C \end{equation} and \begin{equation} \delta_1 - \delta_3 \cong 25^{0}, \delta_3\approx 0. \end{equation} Then \begin{equation} A(D^0\to K^-\pi^+)\cong (2) e^{i ~54.3^{0}} \end{equation} \begin{equation} A(D^0\to K^+\pi^-)\cong -\theta_c^2 (3.7) e^{i ~41.2^{0}} \end{equation} leading to a phase difference of $\delta=13^{0}$. Omitting the W-exchange term gives a slightly smaller value of $6^{0}$. It should be noted that a $\delta_1=25^{0}$ relates to $\delta_R$ for the $I=1/2$ $0^+$ resonance in the $K\pi$ channel by \begin{equation} \tan \delta_R = {\Gamma \over {2 \Delta}} = {B \sin\delta_1 \over {B \cos\delta_1 + (1- B)}} \end{equation} where $B=BR(0^+\to K \pi)\approx 0.50$, $\Gamma\approx 200$ MeV, $\Delta=M_R- M_D\approx 70$ MeV and $\delta_R = (55-65)^{0}$. The models discussed above predict \begin{equation} { {BR(D^0\to K^+\pi^-)} \over { BR(D^0\to K^-\pi^+)}} = (2.3-3.4)\tan^4\theta_c \end{equation} which is compatible with the CLEO~II measurement. There are also other models for $D$ decays in which a value for the phase difference $\delta$ can be extracted\cite{Kaedin}. Since it is difficult to assign errors to these predictions, we regard $0^{0}-13^{0}$ as a reasonable range for $\delta$. In order to explore the range of $\delta$ in the models, we have calculated the value of $\delta$ omitting the W-exchange term. This corresponds to a dramatic change in the parameters of the models. \begin{table}[htb] \caption{Values of $\delta$ in various phenomenological models of $D$ meson decay.} \label{Tbdelta} \begin{tabular}{ll} & $\delta$ \\ \hline Exact SU(3) limit \cite{wolf2} & $0^{0}$ \\ Chau and Cheng\cite{chaucheng} & $4^{0}$ \\ Chau and Cheng (no W-exchange)\cite{chaucheng} & $5^{0}$ \\ Buccella et al. & $13^{0}$ \\ Buccella et al. (no W-exchange) & $6^{0}$ \\ \end{tabular} \end{table} \subsection{Summary of the Interference Effect from FSI} To summarize, the phenomenological models which have been tuned to agree with the observed branching fractions and the fits to the $D$ meson data give $\delta$ in the range of $5^{0}-13^{0}$. To evaluate the possible experimental consequences, consider the case of maximal destructive interference ($\phi=0$, $\beta=1$), with $\delta=13^{0}$. We allow a one standard deviation variation on $R_{DCSD}=\alpha^2$ from the CLEO~II measurement in order to obtain an upper limit on the effect of the interference term. We set $r_{mix}$, the ratio of integrated rates for mixed events relative to unmixed events, to the E691 upper bound\cite{defrmix}. The contributions of the mixing term, the DCSD term, and the term proportional to $\Delta M t$ are shown in Figure 1. These time dependent searches are most sensitive to excess events from mixing for $t>0.22$ ps $\displaystyle ={\tau_{D^0}\over 2}$, where the combinatorial backgrounds are manageable and where the mixing term is expected to peak. In addition, there is no loss in efficiency for the mixing component when this cut is imposed. An upper limit of $t<4.0$ ps is also imposed. The change in the observed event yield for various values of $R_{DCSD}$ and maximal destructive interference are given in Table~\ref{Tbyield}. These were calculated for the scenario with maximal destructive interference and $\delta=13^{0}$. We also give the change in the observed event yield for $t>2\tau_{D^0}$ (this is the region where mixing peaks and the experiments are most sensitive) in Table~\ref{Tblong}. This change is at most 10-15\% and is well within the experimental systematic error assigned by the E691 and E791 experiments to their limits. \begin{table}[htb] \caption{The change in wrong sign event yield for $t>0.22$ ps with maximal destructive interference, $r_{mix}=0.37\%$, and $\delta=13^{0}$.} \medskip \label{Tbyield} \begin{tabular}{ll} $R_{DCSD}$ & $\Delta$ Yield (\%) \\ \hline 0.0052 & $ 10\%$ \\ 0.0077 & $ 8\%$ \\ 0.0102 & $ 1\%$ \\ \end{tabular} \end{table} \begin{table}[htb] \caption{The change in wrong sign event yield for $t>0.88$ ps (2$\tau_{D^0}$) with maximal destructive interference, $r_{mix}=0.37\%$, and $\delta=13^{0}$.} \medskip \label{Tblong} \begin{tabular}{ll} $R_{DCSD}$ & $\Delta$ Yield (\%) \\ \hline 0.0052 & $ 12\%$ \\ 0.0077 & $ 10\%$ \\ 0.0102 & $ 9\%$ \\ \end{tabular} \end{table} \section{Effects of CP Violation} Now consider the contribution of CP violation. Let $\beta=1 -\epsilon$ and $\displaystyle{1\over \beta}=1+\epsilon$\cite{epsfoot}. We assume $\epsilon$ is small compared to 1 and retain only terms linear in $\epsilon$; this is justified in the SM and even more so when $\Delta M$ is enhanced and $\Delta \Gamma/\Delta M<<1$. We allow the phase $\phi$ to be arbitrary. With these definitions and $\Delta \Gamma < <\Delta M$, the expression for $\Gamma(D^0(t)+\bar{D}^0(t))$ now becomes: \begin{eqnarray} \Gamma(D^0(t)+\bar{D}^0(t)) & = & \\ 2 F(t) \times & & \{ 4 \alpha^2 + ({\Delta M}^2 + {{\Delta \Gamma}^2 \over 4}) t^2 \nonumber \\ & & + \alpha (\cos(-\phi+\delta)+\cos(\phi+\delta))\Delta \Gamma ~t +\alpha \epsilon (\cos(\phi+\delta)- \cos(-\phi+\delta))\Delta \Gamma ~t \nonumber \\ & & +2 \alpha (\sin(-\phi+\delta) + \sin(\phi+\delta))\Delta M ~t +2 \alpha \epsilon(\sin(-\phi+\delta) - \sin(\phi+\delta))\Delta M ~t \} \nonumber \end{eqnarray} The quantity $\epsilon$ is assumed to be small as in Ref\cite{Blaylock}, however, the CP violating phase $\phi$ can be large as is the case for certain extensions of the Standard Model. The quantity $\epsilon$ for $D$ mixing is given by\cite{BGHP} \begin{equation} \epsilon \approx {-2 ~{\rm Im}({M_{12}^* \Gamma_{12} \over 2}) \over { {1\over 2} \Delta M^2 + {\Delta \Gamma^2\over 4} } } \end{equation} in the Standard Model and is already small ($\epsilon<O(2\%)$)\cite{BGHP}. In new physics scenarios with $\Delta\Gamma_{SM}=\Delta\Gamma<<\Delta M$, \begin{equation} \tan\phi \cong {{\rm Im}(M_{12}) \over \Delta M} \end{equation} For non standard models with ${\rm Im}(M_{12})/\Delta M$ of order unity, $\tan\phi$ may be large (O(1)). By contrast, \begin{equation} {\epsilon} \cong 2 ( {\Delta \Gamma \over \Delta M}) {{\rm Im}(M_{12})\over \Delta M} \approx 2({\Delta\Gamma \over \Delta M}) << 1 \end{equation} The crucial point is that $\epsilon$ is proportional to $1/\Delta M$ and is highly suppressed if $\tan\phi$ is of order unity and $\Delta M$ is enhanced. It is important to note that while $\tan(\phi)$ can be much larger than the Standard Model expectation $\epsilon$ will be even smaller than the value in the Standard Model for new physics scenarios in which $\Delta M$ is enhanced. The total wrong sign rate can then be reduced to \begin{eqnarray} \Gamma(D^0(t)+\bar{D}^0(t)) & = & 2 F(t) \\ & & \{ 4 \alpha^2 + ({\Delta M}^2 + {{\Delta \Gamma}^2 \over 4}) t^2 \nonumber \\ & & + 2 \alpha (\cos(\phi)\cos\delta))\Delta \Gamma ~t - 2 \alpha \epsilon (\sin(\phi)\sin(\delta))\Delta \Gamma ~t \nonumber \\ & & +4 \alpha \epsilon (\cos(\delta)\sin(\phi))\Delta M ~t +4 \alpha (\sin(\delta) \cos(\phi))\Delta M ~t \} \nonumber \end{eqnarray} With $\epsilon$ as given above and $\Delta \Gamma << \Delta M$, the expression for $\Gamma(D^0(t)+\bar{D}^0(t))$ becomes \begin{eqnarray*} \Gamma(D^0(t)+\bar{D}^0(t)) & = & 2 F(t) \\ & & \{ 4 \alpha^2 + ({\Delta M}^2 ) t^2 \nonumber \\ & & +4 \alpha (\sin(\delta) \cos(\phi))\Delta M ~t \} \nonumber \end{eqnarray*} Hence, the term due to CP violation is too small to be observable when $\Delta M$ and $\rm{Im}(M_{12})$ are enhanced. As experimental sensitivity improves and become sensitive to mixing at the level $r_{mix}<10^{-4}$, it is possible that better sensitivity to $D^0-\bar{D}^0$ mixing can be achieved by fitting the time distribution of $\Gamma(D^0\to K^+\pi^-)-\Gamma(\bar{D}^0\to K^-\pi^+)$. This rate, which will henceforth be denoted $\Gamma(D^0(t)-\bar{D}^0(t))$, is given by \begin{eqnarray} \Gamma(D^0(t)-\bar{D}^0(t)) & = & 2 F(t) \times \\ && \{ 2 \alpha \epsilon (\cos(\phi)\cos\delta))\Delta \Gamma ~t - 2 \alpha (\sin(\phi)\sin(\delta))\Delta \Gamma ~t \nonumber \\ && +4 \alpha (\cos(\delta)\sin(\phi))\Delta M ~t +4 \alpha \epsilon (\sin(\delta) \cos(\phi))\Delta M ~t \} \nonumber \end{eqnarray} In the limit that $\Delta\Gamma<< \Delta M$ and $\phi$ is large, this reduces to \begin{equation} \Gamma(D^0(t)-\bar{D}^0(t)) \cong 2 F(t) \{ 4 \alpha (\cos(\delta)\sin(\phi))\Delta M ~t +4 \alpha \epsilon (\sin(\delta) \sin(\phi))\Delta M ~t \} \end{equation} or neglecting the small term proportional to $\epsilon \sin(\delta)$, \begin{equation} \Gamma(D^0(t)-\bar{D}^0(t)) \cong 2 F(t) [4 \alpha (\cos(\delta)\sin(\phi))\Delta M ~t ] \end{equation} Note that in this case, the long lived tail of DCSD does not contribute to the signal. In addition, as noted by Wolfenstein\cite{wolf2}, for small values of $\Delta M$, the term proportional to $\Delta M ~t$ will be larger than the term in $\Gamma(D^0(t)+\bar{D}^0(t))$ which is proportional to $(\Delta M ~t)^2$. This feature is illustrated in Figs. 2~(a),~2~(b). \section{Conclusions} The formalism presented here must be modified for the case of multibody modes such as $D^0\to K^+ \pi^- \pi^0$ or $D^0\to K^+ \pi^- \pi^+ \pi^-$. For these other modes, an additional complication is that the value of the final state phase difference, $\delta$, may be different from the value in the case of $D^0\to K^+ \pi^-/ \bar{D}^0\to K^+\pi^-$ and is not guaranteed to be small. It should also be remembered that limits on $D^0-\bar{D}^0$ mixing from studies of semileptonic decays do not have the complications from DCSD and other hadronic effects discussed here. At the present level of sensitivity and with reasonable (though model dependent) values for the phase difference $\delta$, the $\Delta M~t$ term which arises from FSI does not dramatically change the observed event yield for experiments which study the time dependence of mixing and is not yet a significant systematic experimental limitation. We suggest that future experiments determine systematic errors on their limits by using an upper limit on the phase difference $\delta$. The contribution from the corresponding term proportional to $\Delta M~t$ due to CP violation which arises in extensions of Standard Model is highly suppressed. This term is not observable at the present level of experimental sensitivity. However, as emphasized by Liu\cite{liu1} and by Wolfenstein\cite{wolf2}, this term should not be neglected as experimental examination of the $D^0(t)-\bar{D}^0(t)$ distribution may allow more sensitive searches for $D^0-\bar{D^0}$ mixing in the future if the CP violating phase is large. This work was supported in part by the United States Department of Energy under grant DE-FG 03-94ER40833 and by Tokkuri Tei. We thank G. Burdman, E. Golowich, J. Hewett, D. Kaplan, T. Liu, Y. Nir, and M. Witherell for useful and enjoyable discussions.
\section{Introduction} \label{sec:int} The mechanism for breaking the symmetry of the electroweak interactions has not been directly accessed in experiments thus far. One possibility is that the Higgs boson is so heavy that it will not be produced even in the next generation of colliders. In this case and in other strongly interacting symmetry-breaking scenarios, it is interesting to parametrize the symmetry-breaking sector in a model independent way through the use of chiral lagrangians \cite{appel}. In this approach, the low energy effects of new physics are represented by an infinite tower of non-renormalizable effective operators which are consistent with the $SU(2)_L \otimes U(1)_Y$ symmetry of the standard model (SM). The lowest order chiral Lagrangian exhibits an universal behavior for the dynamics of the electroweak interactions, being independent of the details of the mechanism of symmetry-breaking. However, at the next-to-leading order in the chiral expansion, there are 14 effective operators whose coefficients are dictated by the underlying dynamics. Among the next-to-leading operators, there is only one that is CP conserving but parity violating \cite{apwu}. This operator also breaks the custodial $SU(2)_C$ symmetry \cite{cust} and is given by \begin{equation} {\cal L}_{11} = \alpha_{11} g~ \epsilon^{\alpha \beta \mu \nu} \text{Tr} \left ( \tau^3 U^\dagger D_\mu U \right )~ \text{Tr} \left ( U^\dagger W_{\alpha\beta} D_\nu U \right ) \; , \label{lag:gz5} \end{equation} where the dimensionless unitary unimodular matrix $U = \exp ( i \xi^a \tau^a / v^2 )$ contains the would-be Goldstone bosons $\xi^a$, $v \simeq 246$ GeV is the symmetry-breaking scale, the $SU(2)_L \otimes U(1)_Y$ covariant derivative is \begin{equation} D_\mu U = \partial_\mu U + i~ \frac{g}{2}~ W^j_\mu \tau^j U - i~ \frac{g^\prime}{2}~ B_\mu U \tau^3 \; , \end{equation} and the field strength tensors are written in terms of $W_\mu = W^j_\mu \tau^j$ \begin{eqnarray} W_{\mu\nu} &=& \frac{1}{2} \left ( \partial_\mu W_\nu - \partial_\nu W_\mu + \frac{i}{2} [ W_\mu, W_\nu] \right ) \; , \\ B_{\mu\nu} &=& \frac{1}{2} \left ( \partial_\mu B_\nu - \partial_\nu B_\mu \right ) \tau^3 \; . \end{eqnarray} The physical content of the above operator is more transparent in the unitary gauge, $U = 1$, where the effective Lagrangian (\ref{lag:gz5}) gives rise to anomalous contributions to the triple vertex $W^+ W^- Z$ and to the four-gauge-boson vertex $W^+ W^- Z \gamma$. In the standard notation of Ref.\ \cite{dieter}, we have the correspondence for the triple gauge-boson vertex \begin{equation} g^Z_5 = \frac{e^2}{s_W^2 c_W^2} \alpha_{11} \; , \end{equation} where we denote the sine (cosine) of the weak mixing angle by $s_W$ ($c_W$). The expected size of $g^Z_5$ depends upon whether or not the underlying dynamics respects the custodial symmetry. In models with a custodial symmetry, $g^Z_5$ should be of the order of $10^{-4}$, while for models without this symmetry we expect $g^Z_5 \sim 10^{-2}$. At low energies, the bounds on this operator come from one-loop contributions to meson decays and to the vertex $ Z f \bar{f}$. From the study of the decay $K_L \rightarrow \mu^+ \mu^-$, we obtain limits of the order $g^Z_5 \lesssim 1$ \cite{kl}, while the precise measurements of the $Z$ flavor diagonal couplings imply that $g^Z_5 \lesssim 0.04$ \cite{sdaw}. These bounds are obtained using the naturalness assumption that no cancellations take place between contributions from different anomalous interactions. However, a closer look at the interaction (\ref{lag:gz5}) reveals that it is momentum dependent, and consequently it can be better studied directly in processes at high energies. The Next Linear $e^+e^-$ Collider (NLC) \cite{pal} will reach a center-of-mass energy between 500 and 2000 GeV with an yearly integrated luminosity of at least $10$ fb$^{-1}$. An interesting feature of this new machine is the possibility of transforming an electron beam into a photon one through the laser backscattering mechanism \cite{las0,laser}. This process will allow the NLC to operate in three different modes, $e^+e^-$, $e\gamma$, and $\gamma\gamma$, opening up the opportunity for a wider search for new physics. However, it is important to stress that the collider can operate in only one of its three modes at a given time, therefore, it is imperative to study comparatively the different features of each of these setups. Previously, the phenomenological implications of the operator (\ref{lag:gz5}) to the reaction $e^+e^- \rightarrow W^+ W^- Z $ at high energies were analyzed in Ref.\ \cite{sdaw:ee}, which showed that it is possible to obtain limits of the order of $g^Z_5 \lesssim 0.3$ for a center-of-mass energy of 500 GeV. However, this sensitivity to $g^Z_5$ can only be achieved for a high degree of $e^-$ polarization. This interaction was also studied in $e\gamma$ collisions in Ref.\ \cite{sdaw:eg} through the process $e^- \gamma \rightarrow W^- Z \nu_e$, that will be able to lead to constraints $g^Z_5 \lesssim 0.12$ for a center-of-mass energy of 500 GeV and an integrated luminosity of $10$ fb$^{-1}$. It is interesting to notice that the bounds obtained in the above processes originate from the direct tree-level contributions of the anomalous interaction. In this work we examine the capability of the next generation of $e^+e^-$ colliders operating in the $\gamma\gamma$ mode to place direct bounds on the effective operator (\ref{lag:gz5}) through the reaction $\gamma \gamma \rightarrow W^+ W^- Z$ \cite{our,our:ano}. In a $\gamma\gamma$ collider, this process exhibits tree-level contributions from the anomalous interaction (\ref{lag:gz5}) and contains the minimum number of final state particles. We show that for a center-of-mass of $500$ ($1000$) GeV it is possible to obtain bounds $g^Z_5 \lesssim 0.15$ ($4.5 \times 10^{-2}$) for an integrated luminosity of $10$ fb$^{-1}$. \section{Results} The most promising mechanism to generate hard photon beams in an $e^+ e^-$ linear collider is laser backscattering. Assuming unpolarized electron and laser beams, the backscattered photon distribution function \cite{laser} is \begin{equation} F_{\gamma/e} (x,\xi) \equiv \frac{1}{\sigma_c} \frac{d\sigma_c}{dx} = \frac{1}{D(\xi)} \left[ 1 - x + \frac{1}{1-x} - \frac{4x}{\xi (1-x)} + \frac{4 x^2}{\xi^2 (1-x)^2} \right] \; , \label{f:l} \end{equation} with \begin{equation} D(\xi) = \left(1 - \frac{4}{\xi} - \frac{8}{\xi^2} \right) \ln (1 + \xi) + \frac{1}{2} + \frac{8}{\xi} - \frac{1}{2(1 + \xi)^2} \; , \end{equation} where $\sigma_c$ is the Compton cross section, $\xi \simeq 4 E\omega_0/m_e^2$, $m_e$ and $E$ are the electron mass and energy respectively, and $\omega_0$ is the laser-photon energy. The quantity $x$ stands for the ratio between the scattered photon and initial electron energy and its maximum value is \begin{equation} x_{\text{max}}= \frac{\xi}{1+\xi} \; . \end{equation} In what follows, we assume that the laser frequency is such that $\xi = 2(1 +\sqrt{2})$, which leads to the hardest possible spectrum of photons with a large luminosity. The cross section for $W^+W^-Z$ production via $\gamma\gamma$ fusion can be obtained by folding the elementary cross section for the subprocesses $\gamma\gamma \rightarrow W^+W^-Z$ with the photon-photon luminosity ($dL_{\gamma\gamma}/dz$), {\it i.e.}, \begin{equation} d\sigma (e^+e^-\rightarrow \gamma\gamma \rightarrow WWZ)(s) = \int_{z_{\text{min}}}^{z_{\text{max}}} dz ~ \frac{dL_{\gamma\gamma}}{dz} ~ d \hat\sigma (\gamma\gamma \rightarrow WWZ) (\hat s=z^2 s) \; , \end{equation} where $\sqrt{s}$ ($\sqrt{\hat{s}}$) is the $e^+e^-$ ($\gamma\gamma$) center-of-mass energy, $z^2= \tau \equiv \hat{s}/s$, and the photon-photon luminosity is \begin{equation} \frac{d L_{\gamma\gamma}}{dz} = 2 ~ z ~ \int_{z^2/x_{\text{max}}}^{x_{\text{max}}} \frac{dx}{x} F_{\gamma/e} (x,\xi)F_{\gamma/e} (z^2/x,\xi) \; . \label{lum} \end{equation} The analytical calculation of the cross section for the subprocess $\gamma\gamma \rightarrow W^+W^-Z$ requires the evaluation of 12 Feynman diagrams in the unitary gauge and it is very lengthy and tedious despite being straightforward. We evaluated numerically the helicity amplitudes for this process using the techniques outlined in Refs.\ \cite{bar:num,zep:num} in order to obtain our results in an efficient and reliable way. As a check of our results, we explicitly verified that the amplitudes were Lorentz and $U(1)_{\text em}$ invariant. The phase space integrations were performed numerically using the Monte Carlo routine VEGAS \cite{lepage}. The total cross section for the process $\gamma \gamma \rightarrow W^+ W^- Z$ is a quadratic function of the anomalous coupling $g^Z_5$, {\it i.e.} \begin{equation} \sigma_{\text{tot}} = \sigma_{\text{sm}} + g^Z_5~ \; \sigma_{\text{int}} + (g^Z_5)^2~ \; \sigma_{\text{ano}} \; , \label{base} \end{equation} where $\sigma_{\text{sm}}$ stands for the SM cross section \cite{our} and $\sigma_{\text{int}}$ ($\sigma_{\text{ano}}$) is the interference (pure anomalous) contribution. We evaluated these contributions for unpolarized backscattered photons imposing that the polar angles of the produced vector bosons with the beam pipe are larger than $10^\circ$. In Table \ref{sigmas:z}, we present our results for several $e^+e^-$ center-of-mass energies. The interference term vanishes since the anomalous amplitude has a phase of $90^\circ$ with respect to the standard model amplitude for an unpolarized initial state. In order to quantify the effect of the new couplings, we defined the statistical significance ${\cal S}$ of the anomalous signal \begin{equation} {\cal S} = \frac{|\sigma_{\text tot} - \sigma_{\text sm}|}{\sqrt{\sigma_{\text sm}}} \; \sqrt{\cal L} \; , \label{sig} \end{equation} which can be easily evaluated using the parametrization (\ref{base}) with the coefficients given in Table \ref{sigmas:z}. We list in Table \ref{const} the values of the anomalous couplings that correspond to a $3\sigma$ effect in the total cross section for the different center-of-mass energies of the associated $e^+e^-$ collider, assuming an integrated luminosity ${\cal L}= 10$ fb$^{-1}$. From this table, we can learn that a $\gamma\gamma$ collider leads to bounds on $g^Z_5$ that are better than the ones that can be obtained in the usual $e^+e^-$ mode. Moreover, $\gamma\gamma$ and $e\gamma$ collider lead to similar constraints on $g^Z_5$. The kinematical distributions of the final state particles can be used, at least in principle, to increase the sensitivity of the $\gamma\gamma$ reactions to anomalous interactions, improving consequently the bounds on them. In order to reach a better understanding of the effects of the anomalous interaction (\ref{lag:gz5}) in the reaction $\gamma \gamma \rightarrow W^+ W^- Z$, we present in Fig.\ \ref{fig:1}--\ref{fig:3} various representative distributions of the final state gauge bosons, adopting the values of the anomalous coupling constants that lead to a $3\sigma$ deviation in the total cross section. In Fig.\ \ref{fig:1} we show the normalized distribution in the rapidity $y_W$ of the $W^\pm$ for a center-of-mass energy of $0.5$ and $1$ TeV. The distributions for $W^+$ and $W^-$ coincide due to the absence of the interference term in the cross section. It is interesting to notice that the anomalous coupling $g^Z_5$ enhances the production of $W^\pm$ in the central region of the detector, where they can be more easily reconstructed. Furthermore, increasing the center of mass energy, the $W$'s tend to populate the high rapidity region, as it happens in the process $\gamma\gamma \rightarrow W^+ W^-$. Consequently, the cut in the $W$ angle with beam pipe discards a larger fraction of events at high energies. The normalized invariant mass distributions of $W^\pm Z$ pairs are presented in Fig.\ \ref{fig:2} for a center-of-mass energy of $0.5$ and $1$ TeV. Once again the $W^+Z$ and $W^-Z$ curves coincide. From this Figure we can learn that the presence of the anomalous interaction increases slightly the invariant mass of the $W^\pm Z$ pairs since the new couplings are proportional to the photon momentum. Moreover, as the center-of-mass energy of the collider is increased, the distributions broaden and shift toward higher invariant masses. Figure \ref{fig:3} shows the laboratory energy distribution of the $Z$ gauge boson. As we can see from this figure, the introduction of the anomalous interaction favors the production of more energetic $Z$ bosons, because of the new momentum-dependent couplings. At lower center-of-mass energies the distribution is rather peaked around small values for the energy of the $Z$ boson because of the available phase space. However, as the center-of-mass energy of the collider increases, the distributions broaden, exhibiting many $Z$ bosons with high energies. Up to this point we were able to demonstrate that a $\gamma\gamma$ collider can reveal the existence of an anomalous interaction such as the one described by (\ref{lag:gz5}). However, the determination that the anomalous events are because of this interaction is a much harder task. In principle this could be done through the study of kinematical distributions. Notwithstanding, several anomalous interactions lead to distributions similar to the ones that we presented, see for instance Ref.\ \cite{our:ano}. The effective operator (\ref{lag:gz5}) could be singled out through the forward-backward asymmetry associated to parity violation, however, this does not happen for unpolarized photons since the interference term vanishes, see Table \ref{sigmas:z}. Therefore, in order to determine which anomalous interaction is responsible for the anomalous events we must employ polarized backscattered photons. As an illustration, we show in Fig.\ \ref{fig:4} the normalized $W^\pm$ rapidity distributions for the subprocess $\gamma \gamma \rightarrow W^+ W^- Z$ with $\sqrt{\hat{s}} = 0.5$ TeV, assuming that one photon has a left-handed polarization while the other is right-handed. As we can see from this figure, the rapidity distribution for the $W^+$ and $W^-$ do not coincide, despite the result being clearly $CP$ invariant. This a feature unique to the anomalous interaction (\ref{lag:gz5}). \section{Conclusions} We analyzed in this work the capability of an $e^+e^-$ collider operating in the $\gamma\gamma$ mode to unravel the existence of the anomalous interaction (\ref{lag:gz5}). We demonstrated that for a center-of-mass energy of $0.5$ ($1$) TeV and an integrated luminosity of $10$ fb$^{-1}$, the study of the reaction $\gamma\gamma \rightarrow W^+ W^- Z$ can lead to bounds $|g^Z_5| \le 0.15$ ($4.5 \times 10^{-2}$). These bounds are similar to the ones that can be obtained in the $e\gamma$ mode of the collider and are better than the one steaming from the usual $e^+e^-$ mode. Moreover, at higher energies the luminosity of $\gamma\gamma$ colliders can be larger than the corresponding $e\gamma$ because of problems in the construction this last mode \cite{berk}. Consequently, the $\gamma\gamma$ mode will be the most powerful one to analyze the $g^Z_5$ anomalous coupling. \acknowledgments This work was partially supported by the U.S. Department of Energy under Contract No. DE-FG02-95ER40896, by the University of Wisconsin Research Committee with funds granted by the Wisconsin Alumni Research Foundation, by Conselho Nacional de Desenvolvimento Cient\'{\i}fico e Tecnol\'ogico (CNPq), and by Funda\c{c}\~ao de Amparo \`a Pesquisa do Estado de S\~ao Paulo (FAPESP).
\section{Introduction} The Calogero-Sutherland (CS) models \cite{Calogero,CS,Moser} describe one-dimensional quantum many-body systems with inverse-square long-rang interactions. Among many variants of the CS model \cite{rev-CS}, a class of models which are not translationally invariant has been known over the passed years \cite{O-Pa}. In particular the so-called CS model of $BC_N$-type (abbreviated as the $BC_N$-CS model hereafter) is the most general model with $N$ interacting particles. The $BC_N$-CS model is intimately related to the root system of type $BC_N$ and invariant under the action of the Weyl group of type $B_N$. Namely, the model is invariant under coordinate transformations \begin{equation} \label{Weyl-group} (q_1,q_2,\cdots,q_N) \mapsto (\epsilon_1 q_{\sigma(1)}, \epsilon_2 q_{\sigma(2)},\cdots, \epsilon_N q_{\sigma(N)}), \end{equation} where $(q_1,q_2,\cdots,q_N)\in \mbox{{\bf R}}^N$ denote the coordinates of $N$ particles, $\epsilon_j\in\{\pm 1\}$ and $\sigma$ is an element of the symmetric group of $N$ letters. Roughly speaking, the Weyl group of type $B_N$ consists of the ordinary exchange of particle coordinates and the sign change of coordinates. As we will see below the latter is understood as the mirror image of particles with respect to a boundary. Recent works have made it clear that the $BC_N$-CS model is relevant to one-dimensional physics with boundaries. For instance, it was pointed out that the non-relativistic dynamics of quantum sine-Gordon solitons in the presence of a boundary is described by the $BC_N$-CS model (with $\sinh$-interaction)\cite{K-S94}. This model is interesting in view of the quantum electric transport in mesoscopic systems\cite{B-R,Caselle}. The Haldane-Shastry model, which is the discrete version of the CS model, with open boundary conditions can also be constructed by utilizing the root system of type $BC_N$ \cite{S-A94,B-P-S}. We shall present further evidence for the relevance of the $BC_N$-CS model to our understanding in one-dimensional physics including boundary effects. In this article we will analyze the long-distance critical properties of the $BC_N$-CS model. Since the exact energy spectrum of the model is available \cite{Nao94}, we may apply the method of finite-size scaling developed in conformal field theory (CFT) to study the critical behavior. The same technique has already been employed when the critical properties of the CS model of $A_{N-1}$-type were considered \cite{K-Y91}. The universality class of the $A_{N-1}$-CS model is identified as a Tomonaga-Luttinger liquid which is equivalent to $c=1$ Gaussian CFT. In what follows we will show that, in contrast to the $A_{N-1}$-CS model, the $BC_N$-CS model exhibits the critical behavior described by $c=1$ CFT with boundaries \cite{Cardy}. Hence the universality class will be found to be a chiral Tomonaga-Luttinger liquid \cite{ch-TLL}. In the next section we first introduce the $BC_N$-CS model and review the energy spectrum of the model obtained by using the asymptotic Bethe-ansatz. In section 3 we consider the thermodynamic properties. In section 4 the finite-size scaling analysis of the energy spectrum is performed. Finally, in section 5, we discuss various critical exponents of correlation functions. \section{The $BC_N$-CS model} Let us write down the Hamiltonian of the $BC_N$-CS model \cite{O-Pa}. We put the system in finite geometry with linear size $L$ and impose periodic boundary conditions. The Hamiltonian is then given by \begin{eqnarray} {\cal H} = -\sum_{j=1}^N \frac{\partial^2}{\partial q_j^2} &+& 2\lambda(\lambda-1) \left(\frac{\pi}{L}\right)^2 \sum_{1\leq j<k\leq N} \left\{ \frac{1}{\displaystyle{ \sin^2\frac{\pi}{L}(q_j-q_k)}} +\frac{1}{\displaystyle{ \sin^2\frac{\pi}{L}(q_j+q_k)}} \right\} \nonumber \\ + \lambda_1(\lambda_1 &+& 2\lambda_2-1) \left(\frac{\pi}{L}\right)^2 \sum_{j=1}^N \frac{1}{\displaystyle{ \sin^2\frac{\pi}{L} q_j}} + 4\lambda_2(\lambda_2-1) \left(\frac{\pi}{L}\right)^2 \sum_{j=1}^N \frac{1}{\displaystyle{ \sin^2\frac{\pi}{L} 2q_j}}, \label{b-tri-hamiltonian1} \end{eqnarray} where $\lambda, \, \lambda_1$ and $\lambda_2$ are coupling constants which are assumed to be non-negative. It is clearly seen that the Hamiltonian (\ref{b-tri-hamiltonian1}) is invariant under the action (\ref{Weyl-group}) of the Weyl group of type $B_N$. There exist several interaction terms which will need explanation. The term $1/\sin^2(\pi/L)(q_j+q_k)$ expresses the two-body interaction between the $j$-th particle and the ``mirror-image'' (we place a mirror at the origin $q=0$) of the $k$-th particle ($j\ne k$). The term $1/\sin^2(\pi/L)q_j^2$ may be interpreted as the potential due to {\it impurity} located at the origin. The term $1/\sin^2(\pi/L)2q_j$ describes the interaction between the $j$-th particle and its own ``mirror-image''. All these terms required by invariance under the action of the Weyl group of type $B_N$ violate translational invariance. Therefore, the total momentum is not a good quantum number for the $BC_N$-CS model. The Hamiltonian (\ref{b-tri-hamiltonian1}) can be cast into another form just by using the elementary identity $\sin 2A=2\sin A\cos A$. One gets \begin{eqnarray} {\cal H} = -\sum_{j=1}^N \frac{\partial^2}{\partial q_j^2} &+& 2\lambda(\lambda-1) \left(\frac{\pi}{L}\right)^2 \sum_{1\leq j<k\leq N} \left\{ \frac{1}{\displaystyle{ \sin^2\frac{\pi}{L}(q_j-q_k)}} +\frac{1}{\displaystyle{ \sin^2\frac{\pi}{L}(q_j+q_k)}} \right\} \nonumber \\ &+& \mu(\mu-1) \left(\frac{\pi}{L}\right)^2 \sum_{j=1}^N \frac{1}{\displaystyle{ \sin^2\frac{\pi}{L} q_j}} + \nu(\nu-1) \left(\frac{\pi}{L}\right)^2 \sum_{j=1}^N \frac{1}{\displaystyle{ \cos^2\frac{\pi}{L} q_j}}, \label{b-tri-hamiltonian2} \end{eqnarray} where $\mu=\lambda_1+\lambda_2$, $\nu=\lambda_2$. In this form of the Hamiltonian the term $1/\sin^2(\pi/L)(q_j+q_k)$ is regarded as the boundary potential as before, while the last two terms in (\ref{b-tri-hamiltonian2}) are regarded as the impurity potentials with the strength determined by $\mu$ and $\nu$ respectively. The Hamiltonian (\ref{b-tri-hamiltonian2}) is suitable for our present considerations. The eigenvalues of the Hamiltonian (\ref{b-tri-hamiltonian2}) of the $BC_N$-CS model have been obtained by one of the authors\cite{Nao94}\footnote{% Precisely speaking, this reference treated the case with $\nu=0$ (the $B_N$-CS model). However, we can easily obtain the formula for the $BC_N$-CS model. The spectrum was also derived in \cite{B-P-S}.}. The energy spectrum so obtained is shown to be reproduced exactly with the use of the asymptotic Bethe-ansatz (ABA) method \cite{Nao94}. Let us recall the ABA formula for the $BC_N$-CS model. First of all the total energy of the system takes the form \begin{equation} \label{one-exc-ene} E_N = \sum_{j=1}^N {k_j}^2, \end{equation} where pseudomomenta $k_j$'s satisfy $k_1>k_2>\cdots>k_N>0$ and obey the ABA equations \begin{eqnarray} \label{exci-rapidi} k_j L &=& 2\pi I_j + \pi (\lambda-1) \sum_{l=1,l\ne j}^N \left\{ \mbox{sgn}(k_j-k_l) + \mbox{sgn}(k_j+k_l) \right\} \nonumber \\ & & + \pi (\mu-1) \mbox{sgn}(k_j) + \pi (\nu-1) \mbox{sgn}(k_j),\hskip10mm j=1,\cdots,N, \end{eqnarray} with $\mbox{sgn}(x)=1$ for $x>0$, $=0$ for $x=0$ and $=-1$ for $x<0$. Here $I_j\ (j=1,\cdots,N)$ are positive integers with $I_1>I_2>\cdots>I_N>0$. These are quantum numbers which characterize the excited states. We emphasize here that, in contrast to the $A_{N-1}$-CS model, the Fermi surface of the $BC_N$-CS model consists of a single point. This is due to the fact that pseudomomenta $k_j$ which are solutions to (\ref{exci-rapidi}) are distributed only over the semi-infinite region as is shown in Fig.1a. Therefore, in view of the bosonization picture, it implies that the low-energy critical behavior of the $BC_N$-CS model will be effectively described by a left (or right)-moving sector of CFT (see Fig.1b). In addition to this, we also notice that the form of our Bethe-ansatz equations (\ref{exci-rapidi}) is quite close to that appeared in the studies of the nonlinear Schr\"{o}dinger equation on the half line\cite{gaudin71,B-M88} as well as the $XXZ$ model with open boundary conditions \cite{H-Q-B87,A-B-B-B-Q87}. The critical behavior observed in these models \cite{gaudin71,B-M88,H-Q-B87,A-B-B-B-Q87} is well described by boundary CFT \cite{Cardy}. It is inferred from these points that boundary CFT will play a role in our study of the $BC_N$-CS model. {}Finally we rewrite our ABA equation (\ref{exci-rapidi}) for further convenience. As has already been mentioned, all the pseudomomenta $k_j$ are positive. However, one can make a trick so that $k_j$ takes values in $(-\infty, \infty)$ as in the bulk system. To realize this let us define $I_{-j}=-I_j,\, I_0=0,\, k_{-j}=-k_j$ and $k_0=0$ with $j=1,\cdots,N$, then we have \begin{eqnarray} \label{extent} k_j &=& 4\pi\frac{1}{2L}I_j + 2\pi(\lambda-1) \frac{1}{2L} \sum_{l=-N}^N \mbox{sgn}(k_j-k_l) \nonumber \\ & & + \frac{\pi}{L}(\mu+\nu-2)\mbox{sgn}(k_j) - \frac{\pi}{L}(\lambda-1)\mbox{sgn}(2k_j) - \frac{\pi}{L}(\lambda-1)\mbox{sgn}(k_j), \end{eqnarray} where $j=-N,-N+1,\cdots,N$. The last two terms in (\ref{extent}) arise since the summation in (\ref{exci-rapidi}) does not include the terms $l=j$ and $l=0$. Now the system turns out to have linear size $2L$ and the number of particles becomes $2N+1$. Note that the density of the system does not change. This doubling trick is known to be efficient when studying one-dimensional physics with boundaries \cite{gaudin71,B-M88,H-Q-B87,A-B-B-B-Q87}. \section{Thermodynamic Properties} The purpose in this section is to discuss thermodynamics of the $BC_N$-CS model. Let us first consider the system at zero temperature. All the states inside of the interval $[-k_F,k_F]$ are occupied, where the Fermi momentum $k_F$ is defined as $k_F=\mbox{max}\{k_j\}$. The thermodynamic limit is taken by $2L\rightarrow \infty,\ 2N+1\rightarrow \infty$ with the density $(2N+1)/2L$ fixed. As usual we define the density of states by \begin{equation} \label{d-s} \lim_{L\mapsto \infty} \frac{1}{2L(k_j-k_{j+1})} =\rho(k), \end{equation} and the sum is converted into integral \begin{equation} \label{integral} \frac{1}{2L}\sum_{j=-N}^N (\ ) \mapsto \int_{-k_F}^{k_F}dk\rho(k) (\ ). \end{equation} {}From (\ref{extent}), (\ref{d-s}), (\ref{integral}) and $\frac{d}{d x}\mbox{sgn}(x) =2\delta(x)$, it is shown that \begin{eqnarray} \label{twice-aba-eq} 1 = 4\pi\rho(k) &+& 4\pi(\lambda-1) \int_{-k_F}^{k_F} dk'\delta(k-k')\rho(k') + \frac{2\pi}{L}(\mu+\nu-2\lambda)\delta(k), \end{eqnarray} where the boundary effect manifests itself in the last term ($\sim 1/L$). Notice that even for $\mu=\nu=0$, it still modifies the equation. Upon taking the thermodynamic limit one can neglect the boundary term. The resulting equation is the same as for the $A_{N-1}$-CS model \cite{CS}. Then it is immediate to get \begin{eqnarray} \label{density} \rho(k) &=& \frac{1}{4\pi \lambda}, \\ \label{fermi-mo} k_F &=& 2\pi\lambda d, \end{eqnarray} where we have put $d=N/L$. It is also straightforward to compute the ground-state energy, \begin{eqnarray} \label{energy-aba} E^{(0)} = \sum_{j=-N}^{N}(k_j^{(0)})^2 = 2L \int_{-k_F}^{k_F} dkk^2\rho(k) = 2L\cdot \epsilon^{(0)} \end{eqnarray} with $\epsilon^{(0)}=4 \pi^2\lambda^2d^3/3$ in the $2L \rightarrow \infty$ limit. It is not difficult to extend the above analysis to the finite temperature case. At finite temperatures the pseudomomenta distribute over the infinite region $(-\infty,\infty)$. One finds \begin{eqnarray} \label{twice-aba-eq-finite} 1 = 4\pi(\rho(k)+\rho^h(k)) &+& 4\pi(\lambda-1) \int_{-\infty}^{\infty} dk'\delta(k-k')\rho(k') \nonumber \\ &+& \frac{2\pi}{L}(\mu+\nu-2\lambda)\delta(k), \end{eqnarray} where $\rho^h(k)$ is the hole density. Let $2L\rightarrow \infty$, then we have \begin{equation} \label{finite-tba} \rho(k)+\frac{1}{\lambda}\rho^h(k) = \frac{1}{4\pi \lambda}. \end{equation}Following now the familiar procedure, we obtain the thermodynamic Bethe-ansatz equation, \begin{equation} \label{tba} \epsilon(k) = k^2-\mu_c + (\lambda-1)T \log \left\{ 1 + \exp \left( - \frac{1}{T} \epsilon(k) \right) \right\}, \end{equation} where $T$ is the temperature, $\mu_c$ is the chemical potential and the energy density $\epsilon(k)$ of particles is defined by \begin{equation} \label{p-p-d} \frac{\rho(k)}{\rho^h(k)} = \exp \left( -\frac{1}{T} \epsilon(k) \right). \end{equation} Performing the low-temperature expansion of the free energy $F(T)$ which is given by \begin{equation} (F(T)-\mu_c(2N+1))/(2L) = - \frac{T}{4\pi}\int_{-\infty}^\infty dk \log(1+e^{-\frac{1}{T}\epsilon(k)}), \end{equation} we have \begin{equation} \label{lte} F(T) \simeq F(T=0) - \frac{\pi T^2}{6(4\pi \lambda d)}. \end{equation} The second term in (\ref{lte}) is responsible for the linear specific heat $C$ as $T\rightarrow 0$. It is well recognized that the coefficient in $C$ is universal modulo the Fermi velocity $v_{\mbox{{\tiny F}}}$ which is not universal \cite{B-C-N}. In translationally invariant systems the Fermi velocity is determined by the dispersion relation. In the $BC_N$-CS model, however, one cannot rely on the dispersion relation since the momentum is not a good quantum number. So, in order to determine $v_{\mbox{{\tiny F}}}$, we have to take another point of view. As we observed, eqs.(\ref{density}), (\ref{fermi-mo}) and (\ref{finite-tba}) coincide with those obtained in the $A_{N-1}$-CS model. Hence we may regard the $A_{N-1}$-CS model as the bulk counterpart of the $BC_N$-CS model. Since the $A_{N-1}$-CS model is described in terms of $c=1$ CFT \cite{K-Y91} we assume that the central charge for the $BC_N$-CS model is also given by $c=1$. Then, comparing $C$ obtained from (\ref{lte}) to the formula $C=\pi cT/(3v_{\mbox{{\tiny F}}})$ \cite{B-C-N} with $c=1$ we find $v_{\mbox{{\tiny F}}}=4\pi\lambda d$. We shall see in section 5 that the finite-size spectrum is in fact in accord with $c=1$ CFT. \section{Finite-size scaling analysis} In this section we perform the finite-size scaling analysis of the energy spectrum of the $BC_N$-CS model. To begin with, we summarize several fundamental formulas in boundary CFT \cite{Cardy} which we will need to analyze the energy spectrum. Let us first recapitulate the finite-size scaling form of the ground-state energy predicted by conformal invariance under {\it free boundary conditions} \cite{B-C-N} \begin{equation} E^{(0)} = L\epsilon^{(0)} + 2f - \frac{\pi v_{\mbox{{\tiny F}}}}{24L}c \ , \label{fsscft} \end{equation} where $\epsilon^{(0)}$ and $f$ are, respectively, the bulk limits of the ground-state energy density and the boundary energy, $v_{\mbox{{\tiny F}}}$ is the velocity of the elementary excitations. The Virasoro central charge $c$ which specifies the universality class of the system appears as the universal amplitude of the $1/L$ term in (\ref{fsscft}). {}From the scaling behavior of the excitation energy one can read off the boundary critical exponents $x_b$ \cite{Cardy}. This exponent $x_b$ governs the power-law decay (parallel to the boundary surface) of a two-point function. Suppose a critical system on the half-plane $\{(y,\tau)\in \mbox{{\bf R}}_{\geq 0}\times \mbox{{\bf R}}\}$ with a surface at $y=0$. ($y$ is the perpendicular distance from a point $(y,\tau)$ to the boundary and $\tau$ means the imaginary time.) Let ${\cal O}(y,\tau)$ be a local operator. We consider its two-point correlation function $G(y_1,y_2,\tau)= \langle{\cal O}(y_1,\tau_1){\cal O}(y_2,\tau_2)\rangle$, which is a function of $\tau=\tau_1-\tau_2$ because of translational invariance along the surface. For $|\tau|\gg y_1,\, y_2$, we obtain the asymptotic form of $G$, \begin{equation} \label{long-time-asymp} G(y_1,y_2,\tau) \sim \frac{1}{\tau^{2x_b}}. \end{equation} To evaluate $x_b$ we have to examine the scaling law \begin{equation} E-E^{(0)} = \frac{\pi v_{\mbox{{\tiny F}}}}{L}x_b \label{ex-corr} \end{equation} with $E$ being the excitation energy. It usually happens that the value of $x_b$ is distinct from that of the bulk exponent for certain scaling operator. In terms of CFT, the bulk exponent is expressed as the sum of left and right conformal weights, while the boundary exponent is equal to the left (or right) conformal weight. Let us now turn to the $BC_N$-CS model. It is convenient to manipulate the ABA equations (\ref{exci-rapidi}) directly. We can easily solve (\ref{exci-rapidi}) to obtain \begin{eqnarray} k_j &=& \frac{2\pi}{L} \left[ I_j - \left( N-j+1 \right) \right] + k_j^{(0)}, \hskip10mm j=1,\cdots,N, \end{eqnarray} where \begin{equation} \label{gra-rapidi} k_j^{(0)} = \frac{2\pi}{L} \left[ \lambda(N-j)+\frac{\mu+\nu}{2} \right]. \end{equation} The ground state is thus specified by the quantum numbers $I_j^{(0)}=N-j+1,\ (j=1,\cdots,N)$, from which we get the Fermi point $I_1^{(0)}=N$ and the Fermi momentum $k_F=2\pi\lambda N/L+\pi(\mu+\nu-2\lambda)/L$. The ground-state energy is then obtained as \begin{eqnarray} \label{one-gro-ene} E_N^{(0)} &=& \sum_{j=1}^N \left( k_j^{(0)} \right)^2 \nonumber \\ &=& \left( \frac{2\pi}{L} \right)^2 \left[ \frac{1}{3} \lambda N + \frac{1}{2} \lambda(\mu+\nu-\lambda)N^2 + \frac{1}{12} \left( 3(\mu+\nu-\lambda)^2-\lambda^2 \right)N^3 \right]. \end{eqnarray} We make a power expansion of (\ref{one-gro-ene}) with respect to $1/L$ while keeping the particle density $d=N/L$ fixed. The result reads \begin{equation} \label{fss-gro} E_N^{(0)} = \epsilon^{(0)}L + 2f + \frac{\pi v_{\mbox{{\tiny F}}}}{L} \lambda(\Delta N_b)^2 - \frac{\pi v_{\mbox{{\tiny F}}}}{12L}\lambda, \end{equation} where $f= \pi^2 \lambda (\mu+\nu-\lambda)d^2$ and \begin{eqnarray} \Delta N_b = \frac{\mu+\nu-\lambda}{2\lambda}. \end{eqnarray} In (\ref{fss-gro}) there appear no higher-order terms with $L^{-m}(m\geq 2)$. Note also the symmetric dependence of $f$ and $\Delta N_b$ on $\mu, \, \nu$. There are several points which should be noticed in (\ref{fss-gro}). First of all, besides the thermodynamic energy density $\epsilon^{(0)}$ already computed in (\ref{energy-aba}), one finds the {\it boundary energy} $2f$ in the term of order $L^0$, which is due to the absence of translational invariance in the system. The next order corrections proportional to $1/L$ turn out to provide valuable information on "boundary effects". To see this, let us proceed a bit carefully by having decomposed the $1/L$-contributions into the last two terms in (\ref{fss-gro}). We first recall that the size-dependence of the interaction is inevitably introduced for $1/r^2$ systems, as seen in (\ref{b-tri-hamiltonian2}), when dealing with interacting particles in finite geometry. This gives rise to nonuniversal $1/L$-corrections to the ground-state energy in addition to the universal one, as observed in the $A_{N-1}$-CS model \cite{K-Y91}. In (\ref{fss-gro}), therefore, we think that the term $-\pi v_{\mbox{{\tiny F}}} \lambda /(12L)$ suffers from such nonuniversal contaminations which, in direct comparison with (\ref{fsscft}), yield the wrong value for the central charge. The other $1/L$-correction term, $\pi v_{\mbox{{\tiny F}}} \lambda (\Delta N_b)^2/L$, is more interesting and understood as the "boundary effect" which consists of two kinds of contributions. As seen from (\ref{extent}), when we convert the $BC_N$ system to the {chiral} system by using a trick of mirror image, we are left with particles moving only in one direction feeling the {\it boundary potential} depending on $\lambda$, in addition to the {\it impurity potential} depending on $\mu$ and $\nu$. These two types of scattering effects are combined into a quadratic form with respect to the "fractional quantum number" $\Delta N_b$ depending on both $\mu+\nu$ and $\lambda$. Note that the quantum number $\Delta N_b$ physically represents the phase shift due to the scattering by the impurity- and boundary-potentials. Thus our ground-state energy $E_N^{(0)}$ is considered as the phase-shifted ground-state energy\cite{A-L94}. If we imagine a hypothetical system which does not include these boundary contributions, the corresponding ground-state energy $\tilde{E}_N^{(0)}$ is written as \begin{equation} \label{new-b-e} \tilde{E}_N^{(0)} = E_N^{(0)} - \frac{2\pi v_{\mbox{{\tiny F}}}}{L} \frac{\lambda}{2}(\Delta N_b)^2. \end{equation} Having discussed the ground-state energy in detail, we next wish to calculate the finite-size corrections to the excited states. Looking at the ABA equations (\ref{exci-rapidi}) let us create an excited state by adding $\Delta N$ particles to the ground-state configuration. In this case, we have the pseudomomenta \begin{equation} \label{p-n-c} k_j= \frac{2\pi}{L} \left[ \lambda(N+\Delta N-j) +\frac{\mu+\nu}{2} \right], \end{equation}from which we immediately obtain the finite-size corrections to leading order in $1/L$, \begin{eqnarray} \label{nun-dis} E_{N+\Delta N}^{(0)}-E_{N}^{(0)} &\simeq& \mu_c^{(0)} \Delta N + \frac{\pi}{L} \left[ 4\pi\lambda(\mu+\nu-\lambda)d\Delta N + 4\pi\lambda^2d(\Delta N)^2\right] \nonumber \\ &=& \mu_c^{(0)} \Delta N + \frac{\pi v_{\mbox{{\tiny F}}}}{L} \lambda \left( \Delta N + \Delta N_b\right)^2 - \frac{\pi v_{\mbox{{\tiny F}}}}{L} \lambda(\Delta N_b)^2, \end{eqnarray} where $\mu_c^{(0)} =\partial \epsilon^{(0)}/\partial d ={k_F}^2$ is the chemical potential. Note that this expression for the finite-size spectrum is essentially the same as that derived for the charge sector in the Kondo problem (see (49) in \cite{fky}). If we redefine $E_N^{(0)}$ by $E_N^{(0)}-\mu_c^{(0)} N$, we find \begin{equation} \label{s-fss} E_{N+\Delta N}^{(0)} - \tilde{E}_N^{(0)} = \frac{2\pi v_{\mbox{{\tiny F}}}}{L} \frac{\lambda}{2} \left(\Delta N + \Delta N_b \right)^2. \end{equation} Since any excitations which carry currents with large momentum transfer are barred due to the absence of translational invariance in the $BC_N$-CS model, the remaining possible type of low-energy excitations are provided by particle-hole excitations labeled by non-negative integers $n$. The corresponding energy is simply obtained by adding $2\pi v_{\mbox{{\tiny F}}}n/L$ to (\ref{s-fss}). Hence we have \begin{equation} \label{s-fss-all} E - \tilde{E}_N^{(0)} = \frac{2\pi v_{\mbox{{\tiny F}}}}{L} \left[ \frac{\lambda}{2} \left(\Delta N + \Delta N_b \right)^2+n \right], \end{equation} where $E$ denotes the energy of the excited state specified by $(\Delta N,\Delta N_b,n)$. In the next section we argue that our result (\ref{s-fss-all}) is in accordance with the scaling law in $c=1$ boundary CFT. \section{Correlation functions} Now that we have evaluated the finite-size corrections it is possible to read off various critical exponents using the scaling relation (\ref{ex-corr}). When comparing our result (\ref{s-fss-all}) with (\ref{ex-corr}) we have to replace $L$ with $2L$ since $L$ has been defined as the periodic length of the system. Bearing this in mind let us take an operator $\psi_b$ which corresponds to the phase-shifted ground state. This operator can be assumed to be the boundary changing operator\cite{A-L94}. With this point of view, the phase-shifted ground state is an excited state relative to $\tilde E^{(0)}_N$ in (\ref{new-b-e}). The scaling dimension of $\psi_b$ is obtained as \begin{equation} x_{\psi_b} = \frac{L}{\pi v_{\mbox{{\tiny F}}}} \left( E_N^{(0)} - \tilde{E}_N^{(0)} \right) = \frac{1} {2\xi^2} \left( \Delta N_b \right)^2, \label{bound-ex-gra} \end{equation} where we have put $\xi=1/\sqrt{\lambda},\ \zeta=1/\sqrt{\mu+\nu}$, and hence $\Delta N_b=(\xi^2-\zeta^2)/(2\zeta^2)$. We next consider an operator $\phi$ which induces the particle number change as well as the particle-hole excitation in the phase-shifted ground state. From (\ref{s-fss}) and (\ref{ex-corr}) we have \begin{equation} x_{\phi} = \frac{L}{\pi v_{\mbox{{\tiny F}}}} \left( E_{N+\Delta N}^{(0)} - \tilde{E}_N^{(0)} \right) = \frac{1} {2\xi^2} \left( \widehat{\Delta N} \right)^2+n, \label{bound-ex} \end{equation} where \begin{equation} \label{modi-q-n} \widehat{\Delta N} = \Delta N + \Delta N_b. \end{equation} Scaling dimensions (\ref{bound-ex-gra}) and (\ref{bound-ex}) take the form of conformal weights characteristic of $c=1$ CFT. The radius $R$ of compactified $c=1$ free boson is taken to be $R=\xi$. Let us concentrate on the self-dual point $R=1/\sqrt{2}$ ({\it i.e.} $\lambda =2$) where the symmetry is known to be enhanced to the level-1 $SU(2)$ Kac-Moody algebra. In the $BC_N$-CS model we have the other continuous parameters $\mu, \, \nu$ which should also be tuned to achieve the $SU(2)$ point. It turns out that $\mu +\nu =0,\, 1,\, 2,\, 3$ and $4$ with $\lambda =2$ are the desired points. This follows from the following observations: When $\mu +\nu =2$ we have $\Delta N_b=0$ and hence \begin{equation} x_{\phi}=\frac{1}{4} (2\Delta N)^2+n \end{equation} which is the conformal weight for the spin-0 irreducible representation of the level-1 $SU(2)$ Kac-Moody algebra. When $\mu +\nu =4$ or $0$ we get $\Delta N_b=\pm 1/2$ and thus \begin{equation} x_{\phi}=\frac{1}{4} (2\Delta N +1)^2+n \end{equation} which is the conformal weight of spin-$1/2$ irreducible representation. When $\mu +\nu =3$ or $1$ we have $\Delta N_b=\pm 1/4$, thereby \begin{equation} x_{\phi}=\frac{1}{16} (4\Delta N +1)^2+n. \end{equation} This is the conformal weight for the unique irreducible representation of the level-1 twisted $SU(2)$ Kac-Moody algebra \cite{twist}. The highest-weight state with $x_\phi =1/16$ is a twist field in $c=1$ CFT. Several $SU(2)$ points identified in \cite{B-P-S} are in agreement with our result. Thus we conclude that the low-energy critical behavior of the $BC_N$-CS model is described in terms of $c=1$ boundary CFT, {\it i.e.} the universality class of a chiral Tomonaga-Luttinger liquid. {}Further considerations on the low-energy critical properties of the $BC_N$-CS model require a clear distinction between two pictures corresponding to two possible sets of quantum numbers. One is a set of quantum numbers $(\Delta N,\Delta N_b,n)$ and the other is a set of $(\widehat{\Delta N},\ n)$ where $\widehat{\Delta N}$ is regarded as the ordinary particle number change in (\ref{bound-ex}) (forgetting about $\Delta N_b$ in (\ref{modi-q-n})). The picture based on the set $(\Delta N,\Delta N_b,n)$ is relevant when describing the long-time asymptotic behavior of the system in which we suddenly turn on the boundary effects in the ground state. The X-ray absorption singularity in the Kondo problem, for instance, is considered in this type of picture \cite{A-L94,fky}. The boundary changing operator $\psi_b$ is described in this picture with $(\Delta N,\Delta N_b,n)=(0,\Delta N_b,0)$. If we use the set $(\widehat{\Delta N},\ n)$ instead, our picture is independent of $\zeta$ and adequate to compute the critical exponents of ordinary correlation functions with boundary effects. Let us consider the one-particle Green function in the above two pictures. Let $(\Delta N,\Delta N_b,n) =(1,\Delta N_b,0)$ in the first picture. This choice of quantum numbers determines the long-time asymptotic behavior of the field correlator (the one-particle Green function) when boundary potentials are turned on at $\tau=0$, \begin{equation} \label{scaling} \langle \Psi^{\dag}(\tau)\Psi(0) \rangle_{{\rm sudden}} \sim \frac{1}{\tau^{2x_{G}}}, \end{equation} where \begin{equation} \label{s-d}x_{G} = \frac{1}{2\xi^2}(1+\Delta N_b)^2 = \frac{1}{8\xi^2} \left( 1 + \frac{\xi^2}{\zeta^2} \right)^2. \end{equation} Here $\langle \cdots \rangle_{{\rm sudden}}$ stands for the expectation value when the boundary potential is suddenly switched on. On the other hand, if we let $(\widehat{\Delta N},n) =(1,0)$ in the second picture, the field correlator takes the form, \begin{equation} \label{scaling2} \langle \Psi^{\dag}(\tau)\Psi(0) \rangle \sim \frac{1}{\tau^{2x_{g}}}, \end{equation} where \begin{equation} \label{s-d2} x_{g} = \frac{1}{2\xi^2}, \end{equation} which describes the ordinary one-particle Green function. In this case, the boundary critical exponent $x_{g}$ linearly depends on $\lambda$. Contrary to these Green functions, the density-density correlation function is controlled by the excitations which do not change the number of particles. Hence, it should have the long-time asymptotic form, \begin{equation} \langle \rho(\tau)\rho(0) \rangle \sim \frac{1}{\tau^{2}}, \end{equation} which follows by taking the quantum number $(\widehat{\Delta N},n)=(0,1)$ in (33). Note that there do not appear anomalous exponents in this correlator. One can easily see that this is also the case for sub-leading terms $\tau^{-2k}$ in which the quantum number is chosen as $(\widehat{\Delta N},n)=(0,k)$. This fact will be confirmed shortly in the following. We now compare our result with the explicit calculations of the dynamical correlation function. In the case $\lambda=1,\, \nu=0$ with $\mu$ arbitrary which corresponds to the noninteracting system, the dynamical density-density correlation function for the $BC_N$-CS model has been obtained by Mac\^{e}do \cite{Macedo} (see also \cite{MIT}). In the thermodynamic limit, the density-density correlation function $G(y_1,y_2,\tau)$ has the form \begin{eqnarray} \label{auto-corr} G(y_1,y_2,\tau) = \frac{\pi^4}{4} y_1 y_2 && \int_1^\infty d u_1 e^{-\frac{1}{2}\pi^2\tau u_1} J_{\mu-\frac{1}{2}}(\pi y_1 \sqrt{u_1}) J_{\mu-\frac{1}{2}}(\pi y_2 \sqrt{u_1}) \nonumber \\ \times && \int_0^1 d u_2 e^{\frac{1}{2}\pi^2 \tau u_2} J_{\mu-\frac{1}{2}}(\pi y_1 \sqrt{u_2}) J_{\mu-\frac{1}{2}}(\pi y_2 \sqrt{u_2}), \end{eqnarray} where $J_{\nu}(z)$ is the Bessel function and $\tau$ is the imaginary time. When $\mu=1/2+m\ (m=0,1,\cdots)$ it is not difficult to evaluate the large-$\tau$ asymptotic behavior by making use of the series expansion of $J_m(z)$. After some algebra we obtain \begin{equation} \label{special-case} G(y_1,y_2,\tau) = \sum_{k=1}^\infty A_k(y_j) \left( \frac{1}{\tau} \right)^{2k} + \sum_{l=0}^\infty B_l(y_j) \left( \frac{1}{\tau} \right)^{l+m+2} e^{-\frac{1}{2}\pi^2\tau}, \end{equation} where $A_k(y_j), B_l(y_j)$ are some functions. As $\tau\rightarrow \infty$ with $y_1,\, y_2$ fixed, the second term vanishes exponentially, yielding \begin{equation} \label{asymp-kappa} G(y_1,y_2,\tau) \simeq \frac{A_1}{\tau^2}+\frac{A_2}{\tau^4} +\frac{A_3}{\tau^6}+\cdots . \end{equation} Notice that the exponents are independent of $m$ ({\it i.e.}, $\mu$). The density-density correlation function is considered in the picture based on $(\widehat{\Delta N},\ n)$. Then we see from (\ref{bound-ex}) that all these exponents are precisely understood in terms of the excitations $(\widehat{\Delta N},\ n)=(0,k)$. This means that the correlation function $G$ is dominated by the particle-hole excitations, and hence there is no way of depending on $\lambda$. Therefore the result (\ref{asymp-kappa}) completely agrees with our prediction by CFT analysis. We are thus led to conclude that the power-law decay in (\ref{asymp-kappa}) is universal irrespective of $\lambda$ (but with $\nu =0$ fixed) though (\ref{asymp-kappa}) is verified at $\lambda =1$. We stress that this remarkable feature in the density-density correlation function is inherent in {\it chiral} Tomonaga-Luttinger liquids \cite{ch-TLL}. {}Finally we briefly mention possible applications to the (chiral) random matrix theory\cite{Mehta91}. Let us recall the $B_N$ Calogero-Moser model ($B_N$-CM model) in the rational form \cite{O-Pa}, \begin{eqnarray} \label{cm-model} {\cal H}_{\mbox{{\tiny C-M}}} = -\sum_{j=1}^N \frac{\partial^2}{\partial x_j^2} &+& 2\lambda(\lambda-1) \sum_{1\leq j<k\leq N} \left\{ \frac{1}{(x_j-x_k)^2} + \frac{1}{(x_j+x_k)^2} \right\} \nonumber \\ &+& \mu(\mu-1) \sum_{j=1}^N \frac{1}{x_j^2} + \omega^2 \sum_{j=1}^N x_j^2, \end{eqnarray} with $\omega>0$. In the thermodynamic limit, this model belongs to the same universality class as the $B_N$-CS model which is equivalent to the $BC_N$-CS model at $\nu =0$. The ground-state wave function for the $B_N$-CM model takes the form of Jastrow-type \cite{O-Pa} \begin{equation} \label{cm-gr-st} \Psi^{(0)} (x_1,x_2,\cdots,x_N) = {\cal N} \prod_{1\leq j<k\leq N} |x_j^2-x_k^2|^{\lambda} \prod_{l=1}^N |x_l^2|^{\frac{\mu}{2}} \exp \left( -\frac{1}{2}\omega x_l^2 \right), \end{equation} where ${\cal N}$ is a calculable normalization constant. Notice that $\Psi^{(0)} (x_1,x_2,\cdots,x_N)$ depends only on the $x_j^2$'s. Then, introducing new variables $z_j=x_j^2$, one should note that $|\Psi^{(0)} (x_1,x_2,\cdots,x_N)|^2$ is identical to the probability distribution function for the eigenvalues $z_j$ of the Laguerre ensemble when $\lambda=1/2,1$ and $2$ (with appropriate values of $\mu$ and $\omega$) corresponding to the ensembles of orthogonal, unitary and symplectic types \cite{Mehta91}, respectively. Therefore, it will be very interesting if the long-time asymptotic behavior of correlation functions in the $B_N$-CM model obtained in the present work is directly compared with the results in the Laguerre random matrix theory. In summary, we have investigated boundary critical phenomena in the $BC_N$-CS model. The boundary effects come from both the impurity potentials and interactions between particles and ``image'' particles. Making use of boundary CFT, we have obtained boundary critical exponents, and clarified the critical properties of the $BC_N$-CS model in terms of chiral Tomonaga-Luttinger liquids. \vskip10mm {\bf Acknowledgements}\\ T.Y. was supported by the Yukawa memorial foundation and the COE (Center of Excellence) researchers program of the Ministry of Education, Science and Culture, Japan. N. K. was partly supported by a Grant-in-Aid from the Ministry of Education, Science and Culture, Japan. The work of S.-K.Y. was supported in part by Grant-in-Aid for Scientific Research on Priority Area 231 ``Infinite Analysis'', the Ministry of Education, Science and Culture, Japan.
\section{Introduction} Experiments on spinodal decomposition in polymer blends show that the coarsening process may sometimes slow dramatically or even cease before reaching equilibrium \cite{hasegawa88} - \cite{lauger94}. In these systems, spinodal decomposition --- which is the process by which a thermodynamically unstable mixture demixes to a stable, phase-separated equilibrium state \cite{cahn} - \cite{glotzer95} --- proceeds normally for some time following a quench to the unstable region ($T<T_s$), and then stops. A break-up of the characteristic, interconnected pattern is observed to precede this pinning phenomenon. The nonequilibrium, microphase-separated blend has been observed to remain in this pinned state over an appreciable time scale where little domain growth occurs. The eventual breakup of the evolving morphology into separated droplets is a natural consequence of the asymmetric composition in an off-critical blend \cite{hayward87}; nevertheless, it may also occur in near-critical blends due to other forces (e.g. gravity). Polymer blends in which pinning has recently been observed for off-critical composition include X-7G/poly(ethylene teraphthalate) (a liquid crystalline polymer/homopolymer blend) \cite{hasegawa88}, poly(styrene-ran-butadiene)/polybutadiene \cite{hashimoto92}, poly(styrene-ran-butadiene)/polyisoprene \cite{hashimoto92}, and polybutadiene/polyisoprene \cite{lauger94}. The specific mechanism responsible for pinning in these blends is poorly understood, and is currently a topic of considerable discussion. While there is general agreement that growth stops soon after the breakup into separated ``droplets'' or ``clusters'' (a so-called ``percolation-to-cluster transition'' \cite{hashimoto92,hayward87}), the mechanism that prevents further coarsening of disconnected domains remains to be clarified. One intriguing scenario points to an entropic barrier as the reason for the observed arrested growth of off-critical phase separating blends. Kotnis and Muthukumar (KM) \cite{kotnis92} have suggested that due to the connectivity of the chains and the reduced conformational entropy near domain interfaces \cite{helfand}, the usual evaporation-condensation mechanism of coarsening \cite{lifshitz61} observed in small-molecule mixtures is suppressed in polymer blends, and instead coarsening occurs via parallel transport of chains along the interface \cite{huse86}. Consequently, KM postulate that if the concentration of the minority-rich phase becomes smaller than the percolation threshold, the parallel coarsening mechanism will be inhibited and the clusters will ``freeze'' after an initial growth period. Hashimoto and coworkers have instead postulated that the enthalpy of mixing, rather than the entropy, provides the barrier to further coarsening following the percolation-to-cluster transition \cite{hashimoto92}. They argue that the increase in enthalpy of mixing suffered upon removing a chain of species $A$ and degree of polymerization $N$ from the surface of an $A$-rich domain is $\Delta H_{\rm mix} \propto \chi N k_BT$, where $\chi$ is the Flory interaction parameter, $k_B$ is Boltzmann's constant, and $T$ is temperature. In the strong segregation limit $\chi N \gg 1$, and thus evaporation of the chain from the domain surface, which would occur with a Boltzmann probability proportional to $\exp (-\Delta H_{\rm mix}/ k_BT)$, is highly unfavorable. Thus, when the parallel transport mechanism is eliminated by the breakup into droplets, coarsening ceases. In this paper, we re-explore the kinetics of spinodal decomposition in off-critical polymer blends described by the Flory-Huggins-De Gennes (FHDG) free energy functional, through numerical simulations of the Cahn-Hilliard (CH) equation. In Sec.~\ref{sec:theory}, we discuss the CH-FHDG equation and the origin of the concentration-dependent square gradient coefficient that has been proposed by KM to cause pinning in off-critical blends. The discretization and numerical integration scheme used to solve this equation, and our numerical results, are presented in Sec.~\ref{sec:solution} and discussed in Sec.~\ref{sec:discussion}. Finally, a summary of our main conclusions, and speculations on possible mechanisms of pinning in blends, is discussed in Sec.~\ref{sec:conclusion}. \section{Theoretical Model} \label{sec:theory} Model blends are typically described by the Flory-Huggins-De Gennes free energy functional \cite{degennes80,pincus81,binder83}: \begin{eqnarray} {F\{\phi({\bf r})\} \over k_B T}= \int d^3r \left[{f_{FH}(\phi ({\bf r}))\over k_BT} + \kappa(\phi)(\nabla \phi)^2 \right], \label{eq:fhdg} \end{eqnarray} with \begin{eqnarray} \kappa(\phi) = {\sigma_a ^2 \over 36 \phi} + {\sigma_b^2 \over 36 (1-\phi)} + \chi \lambda ^2, \label{eq:kappa} \end{eqnarray} where $\phi({\bf r})$ is the local concentration of component $A$ (so that 1-$\phi$ is the concentration of component $B$), $\sigma_A$ and $\sigma_B$ are the Kuhn lengths of the two species, $\lambda$ is an effective interaction distance between monomers, and the Flory-Huggins free energy is \cite{flory,huggins} \begin{eqnarray} {{f_{FH}(\phi)}\over k_BT} = {\phi \over N_A} \ln \phi + {(1-\phi) \over N_B} \ln (1-\phi) + \chi \phi (1-\phi), \label{eq:fh} \end{eqnarray} where $N_A$ ($N_B$) is the degree of polymerization of chains $A$ ($B$). Whereas in small molecule mixtures the square gradient coefficient is enthalpic (arising from short range interactions between molecules) and independent of the local concentration \cite{fredrickson92}, De Gennes proposed that the connectivity of polymer molecules in inhomogeneous blends manifests itself through an additional, concentration-dependent contribution to the square gradient coefficient $\kappa(\phi)$ \cite{degennes80}. The expression for $\kappa(\phi)$ in Eq.~\ref{eq:kappa} was derived to be consistent with the random phase approximation result for the inverse structure factor of an incompressible polymer blend \cite{binder83,edwards66,akcasu}, \begin{eqnarray} S^{-1}({\bf q}) = {1\over {N_A\phi_o D(q^2R_A^2)}} + {1\over {N_B(1-\phi_o) D(q^2R_B^2)}} - 2\chi. \end{eqnarray} Here $R_i^2$ is the average square radius of gyration of species $i$, $\phi_o$ is the average value of the concentration and the Debye function is $D(x) = 2[x-1+ e^{-x}]/x^2$, with $x \equiv q^2R_i^2$. In the weak segregation limit, the interfacial width is much larger than the chain dimensions \cite{fredrickson92}, so that the length scales of interest are larger than $R_i$ ($q^2R_i^2 \ll 1$), and thus the Debye function may be approximated by $D^{-1}(x) = 1+x/3+O(x^2)$. The square gradient coefficient in Eq.~\ref{eq:kappa} is then obtained from the coefficient of the $q^2$ term in the Taylor expansion of the inverse structure factor, which is related to the free energy functional by \cite{binder83} $S^{-1}({\bf q}) = (k_BT)^{-1} \delta^2\!F/\delta \phi^2$, where the r.h.s. is evaluated in q-space. Because of the approximations made in the calculation of the square gradient expression, the Flory-Huggins-de Gennes free energy functional describes the physics of blends in the weak segregation limit, and on length scales much larger than the average chain dimension \cite{binder94,fredrickson92,tang92}. In strongly-segregating blends ($\chi N \gg 1$) for which $\chi$ is small but $N \to \infty$, Eq.~2.2 with a different prefactor in the $\phi$-dependent part is typically used \cite{tang92,roe86,mcmullen92}. The local part of the free energy (Eq.~\ref{eq:fh}) has the same Ginzburg-Landau type of double-well structure as small molecules or Ising-like systems \cite{goldenfeld92}. Thus, the only difference between the free energy functionals for the simplest small molecule and polymeric systems arises from the chain connectivity, which is expressed in the FHDG functional through the reduction of the entropic part of the local term, and by the $\phi$-dependent part of the square-gradient coefficient. KM proposed that the entropic contribution to the nonlocal part of the free energy, namely the concentration-dependent square-gradient coefficient, provides the barrier to coarsening which, when combined with the percolation-to-cluster transition, causes pinning of the off-critical phase-separating blend. In the next section, we re-examine the numerical solution of the time evolution of the CH-FHDG equation for both critical and off-critical blends. Specifically, we show that for this model pinning is {\it not observed} in the continuum limit, regardless of the blend composition, although a dynamical exponent slightly smaller than $1/3$ is found. The theoretical description of spinodal decomposition in binary blends is based on the Cahn-Hilliard-Cook equation for the time evolution of the concentration, originally derived for small molecule systems \cite{cahn,cook}: \begin{eqnarray} {\partial \phi \over \partial t} = \nabla \cdot \left(M(\phi) \nabla {\delta F\{\phi\} \over \delta \phi}\right) + \eta({\bf r},t). \label{eq:ch} \end{eqnarray} In this equation, $M(\phi)$ is the mobility, $F\{\phi\}$ is the coarse-grained free energy functional, and $\eta$ is thermal noise. For polymers, the free energy functional is typically taken to be of the Flory-Huggins-De Gennes form in Eq.~\ref{eq:fhdg}, but more general free energy functionals may be included. In the following, we will always consider for simplicity a symmetric blend, for which $N_A = N_B \equiv N$ and $\sigma_A = \sigma_B \equiv \sigma$. In this case, the mobility \begin{eqnarray} M(\phi) = N D \phi (1-\phi) \label{eq:mob} \end{eqnarray} has been proposed \cite{degennes80}, where $D$ is the self-diffusion coefficient. We will also take the effective interaction distance $\lambda$ equal to the Kuhn length $\sigma$. Substituting in Eq.~\ref{eq:ch} the functional derivative of $F$ and the expression for $M$, the time evolution of the concentration is given by: \begin{eqnarray} {\partial \phi({\bf r}, t) \over \partial t} = N D \nabla \left\{ \phi (1-\phi) \nabla \left[ {1 \over N} \ln {\phi \over 1- \phi} + \chi (1- 2\phi) - \right. \right. \cr \cr \left(2 \chi \sigma^2 + {\sigma^2 \over 18 \phi (1-\phi)} \right) \nabla^2 \phi \left. \left. + {(1-2 \phi) \sigma^2 \over 36 \phi^2 (1- \phi)^2} (\nabla \phi)^2 \right] \right\}. \label{eq:chfhdg1} \end{eqnarray} Note that in Eq.~\ref{eq:chfhdg1} we have neglected the thermal noise term. Since we are interested in the presence or absence of pinning due to the FHDG free energy functional alone, and since it has been shown that the presence of thermal noise in the analogue of this equation for small molecule systems does not influence the scaling function or the growth exponent during coarsening\cite{binder90,rogers88}, we will neglect noise in our simulations\cite{ctgm89,brown93}. Eq.~\ref{eq:chfhdg1} can be rescaled so that it depends on dimensionless space and time variables. The transformation, valid only in the unstable region, is the following \cite{kotnis92}: \begin{eqnarray} {\bf x} = {(\chi - \chi_s)^{1/2} \over \sigma} {\bf r} \hspace{1cm} \tau = {D (\chi - \chi_s)^2 \over \sigma^2 \chi_s} t, \end{eqnarray} where $\chi_s = 1/(2N \phi_0 (1-\phi_0))$ gives the spinodal curve and $\phi_0$ is the average value of concentration. (This rescaling differs from the rescaling commonly used in experiments by a simple numerical factor.) After this transformation, Eq.~\ref{eq:chfhdg1} becomes \cite{kotnis92}: \begin{eqnarray} {\partial \phi({\bf x}, \tau) \over \partial \tau} = {1 \over 2 \phi_0 (1-\phi_0)} \nabla \left\{ \phi (1-\phi) \nabla \left[ {\chi_c \over 2(\chi - \chi_s)} \ln {\phi \over 1- \phi} - 2 {\chi \over \chi - \chi_s} \phi - \right. \right. \cr \cr \left. \left. \left(2 \chi \sigma^2 + {\sigma^2 \over 18 \phi (1-\phi)} \right) \nabla^2 \phi + {(1-2 \phi) \sigma^2 \over 36 \phi^2 (1- \phi)^2} (\nabla \phi)^2 \right] \right\}. \label{eq:chfhdg2} \end{eqnarray} \section{Numerical Solution} \label{sec:solution} The solution of this continuum equation (~\ref{eq:chfhdg2}, which was first studied by KM\cite{kotnis92}, describes the time evolution of the concentration field after a quench to $\chi>\chi_s$ in the unstable region \cite{ctgm89,brown93}. The initial condition before the quench, corresponding to high temperature, is given by a uniform field with random fluctuations about its average value $\phi_0$. At early times following the quench, the uniform concentration is unstable with respect to the long wavelength fluctuations arising from the initial condition, and the two components begin to spatially separate. Domains rich in one or the other component form, and then coarsen so as to remove interfaces and minimize the free energy. In small-molecule mixtures described by, e.g., the Ginzburg-Landau free energy functional, these domains coarsen until phase separation is complete regardless of the relative composition and the presence of thermal noise. Our goal in this paper is to determine if the same is true for the FHDG free energy functional \cite{brown93}. In the late stages of decomposition, the system can be characterized by the evolution of the typical size of growing domains. The scaling hypothesis \cite{binder74} states that this length $L$ (as calculated from, e.g., the inverse of the peak position of the structure factor, the inverse of the first moment of the structure factor, the position of the first zero in the pair correlation function, etc.) evolves in time according to \begin{eqnarray} L \sim \tau^{\alpha}. \end{eqnarray} The choice of one particular definition of $L$ is dictated only by convenience. Numerical simulations, experiments, and analytical results strongly support the validity of the scaling hypothesis in small molecule systems, giving in the absence of hydrodynamic forces the value $\alpha = 1/3$ independent of quench depth, relative composition, and the presence of noise\cite{binder90}. While polymers are believed to belong to the same static universality class as small molecules, the situation is less clear with respect to dynamics. To study the kinetics of spinodal decomposition in polymer blends, we numerically integrate Eq.~\ref{eq:chfhdg2} via a finite difference scheme for both time and space variables. The continuous space of position vectors is replaced by $n^3$ sites on a simple cubic lattice with mesh size (lattice spacing) $\Delta x$. The temporal discretization is achieved by replacing the continuous time variable $\tau$ by a series of $m$ discrete time steps of duration $\Delta \tau$. The value of the concentration field at all sites at each time step is then computed by a first-order Euler numerical integration scheme \cite{numrecipes}, described in detail in the Appendix. Although a large time step and mesh size would speed up the computation, the mesh size must be chosen carefully so as to be smaller than the smallest important length scale in the problem at all times --- here the interfacial width, which decreases in time until the latest stages of demixing. The size of the time step is in turn limited by the mesh size. A time step that is too large could generate instabilities and spurious solutions \cite{rogers88,numrecipes}. Thus, these discrete variables must be chosen carefully in concert. A linear stability analysis can be helpful in suggesting reasonable trial values. If the algorithm is stable with these values of $\Delta x$ and $\Delta \tau$, one can then vary them to find optimum values and to ensure that the solution is accurate and {\it independent} of the choice of these parameters. We have studied the effect on the numerical solution of Eq.~\ref{eq:chfhdg2} of changing $\Delta x$; $\Delta \tau$ is changed suitably so as to maintain stability. The boundary conditions are periodic in all three directions. Initial conditions are given by random values of the concentration field, with average $\phi_0$ and a flat distribution between $\phi_0 - \Delta \phi_0$ and $\phi_0 + \Delta \phi_0$. $\Delta \phi_0$ has a strong influence on the behavior during the initial regime, but does not affect the late stage behavior. Therefore we fix $\Delta \phi_0 = 0.1$. Several realizations of the initial conditions were averaged together for every set of parameters. The phase separation is monitored visually in real space, and quantitatively by determining the time evolution of $k_1(\tau)$, the first moment of the spherically-averaged structure factor \cite{glotzer95}; this is the inverse length that is used to determine the exponent $\alpha$. For the purposes of comparison with previous studies\cite{kotnis92,brown93}, we take $\chi$ to be related to $T$ (K) by \cite{wiltzius}: \begin{eqnarray} \chi = 0.326/T -2.3 \cdot 10^{-4}, \end{eqnarray} and fix $T_c = 62\:^o C$. The system was quenched to temperatures $T=54.5, 49$ and $25 \;^o C$ for critical composition ($\phi_o = 0.5$), and to temperatures $T=35$ and $15 \; ^o C$ for off-critical composition ($\phi_o = 0.4$). We first consider the solution of Eq.~\ref{eq:chfhdg2} obtained with a mesh size $\Delta x = 1$, time step $\Delta \tau = 0.01$ and $n = 32$ (Fig.~\ref{n=32}). This choice of mesh size gives, e.g. for $T=35 \; ^o C$ and $\phi_0 = 0.4$, an equivalent dimensional mesh size of $\Delta r \approx 6 R_g$, where $R_g$ is the average chain radius of gyration. With these choices, we are exactly repeating the integration of Eq.~\ref{eq:chfhdg2} previously performed by Kotnis and Muthukumar. Our results reproduce their findings. For critical quenches ($\phi_0= 0.5$), after an initial transient, the system enters the late stage regime, where $k_1$ decays in time with a power law. The exponent $\alpha$ appears to be smaller than $1/3$, the value expected for spinodal decomposition in small molecule systems and for polymer blends in the intermediate stages of demixing (ie.~without hydrodynamics). For off-critical quenches ($\phi_0=0.4$), domain growth stops before the phase separation is complete. Results change drastically when the mesh size is reduced to $\Delta x = 0.5$, for which the time step must be reduced to $\Delta \tau = 0.002$ to maintain numerical stability (Fig.~\ref{n=64}). (Note that in this case we must take $n = 64$ to keep the system size the same as before --- $n \cdot \Delta x = 32$). This choice of mesh size gives, e.g. for $T=35 \;^o C$ and $\phi_0 = 0.4$, an equivalent dimensional mesh size of $\Delta r \approx 3 R_g$. In this case, the late stage behavior is the same for both critical and off-critical quenches, even after the blend undergoes the percolation-to-cluster transition. After an initial transient, the late stage scaling regime is reached: $k_1$ decays as a power law and no pinning is observed. The values of $\alpha$ can be computed from the slopes of the curves in Fig.~\ref{n=64} and are reported in Table~\ref{tab1}. In all cases $\alpha$ is found to be greater than $1/4$ and smaller than $1/3$. We believe that the latter is the true asymptotic value for this free energy functional; the small systematic error may be attributed to the crossover from the preasymptotic regime and, possibly, to the numerical slowing down given by a still oversized mesh size (see Sec.~\ref{sec:discussion}). Increased accuracy might be obtained by using an even smaller value for $\Delta x$, but this implies a suitable reduction in the value of $\Delta \tau$, and as a result, a prohibitively large computation time. The critical relevance of the mesh size to the late stage behavior of the system is also evident from Fig.~\ref{comparison}, where a ``pinned'' state obtained with $\Delta x = 1$ is ``depinned'' by reducing the mesh size to $0.5$. \section{Discussion} \label{sec:discussion} The absence of pinning for small mesh sizes in our simulations even for off-critical quenches clearly shows that what is observed for $\Delta x = 1$ is not a physical effect but only an artifact of the discretization. By integrating Eq.~\ref{eq:chfhdg2} via a discretization scheme, we are actually changing the model under consideration: the solution of the discrete model exactly reproduces that of the continuum equation only in the limit where $\Delta t$ and $\Delta x$ approach zero. For this reason one must always confirm that the numerical results are independent of the values of the discretization variables. In particular, it was recently shown that an oversized mesh size can cause a non-physical freezing of interfacial motion for systems with non-conserved order parameter\cite{osher94}. ``Spurious pinning'' was already noted by Rogers et al. \cite{rogers88} in a conserved order parameter system, who showed that for the Cahn-Hilliard equation with the Ginzburg-Landau free energy functional, a mesh size $\Delta x >1.7$ causes an unphysical decrease in the effective growth exponent. In general, the solution of the discrete model should reproduce the behavior of the continuum equation if $\Delta x$ is much less than the smallest physical length modeled in the system. In the case under investigation, we are studying a polymer blend in the weak segregation limit ($\chi \ge \chi_c, \chi \ll 1$). The smallest physical length that must be resolved is the interfacial width, which during the late stages of phase separation is of the order of the correlation length $\xi$ \cite{fredrickson92}: \begin{eqnarray} \xi \sim {R_g \over (\chi - \chi_c)^{1/2} \cdot N^{1/2}} \sim {\sigma \over (\chi - \chi_c)^{1/2}}, \end{eqnarray} which is close to unity in rescaled units. Choosing $\Delta x = 1$ implies that we are resolving the system at a length {\it equal} to the correlation length. The spatial derivatives at the interfaces are consequently computed inaccurately with this mesh size, thereby producing unphysical results. This picture is confirmed by Fig.~\ref{profile} showing that when $\Delta x = 1$ the interface appears only one mesh size wide. When $\Delta x = 0.5$ the interface is smoother, and larger than the mesh size; hence no pinning occurs. This effect also explains the low estimates of $\alpha$ for critical quenches when $\Delta x =1$ --- the sharpness of the interface unphysically slows down the evolution of the solution, even if it is not sufficient to pin it. Possibly also the results for $\Delta x = 0.5$ are slightly biased by this effect. It is possible to understand the origin of this problem in another way by looking directly at the equation of motion. Consider for simplicity a one-dimensional small molecule system (i.e. the square-gradient coefficient $\kappa$ and the mobility $M$ are constant), where the concentration profile goes from one bulk value ($\phi_1$) at site $x_i-\Delta x$ to the other ($\phi_2$) at site $x_i+\Delta x$, through an interface. For the domain size to grow, the interface must move, and thus $\phi(x_i)$ must change from $\phi_1$ to $\phi_2$. The driving force for this change is the square gradient term in the free energy, which yields a Laplacian in the functional derivative of $F$. This force must overcome the double well potential given by the local term in the free energy expression, as stated by the Cahn-Hilliard equation: \begin{eqnarray} {\partial \phi \over \partial t} = M \nabla^2 \left({\partial f \over \partial \phi} - \kappa \nabla^2 \phi \right). \end{eqnarray} In the discrete version of this equation, the local part does not depend on $\Delta x$, while the Laplacian is given by: \begin{eqnarray} \nabla^2 \phi = {1 \over (\Delta x) ^2} \left[ \phi(x_i+\Delta x)+ \phi(x_i-\Delta x) - 2 \phi(x_i) \right]. \end{eqnarray} When we increase $\Delta x$ the denominator grows indefinitely, while the numerator is bounded above by $\phi_2+\phi_1- 2 \phi(x_i)= \mbox{const}$. Then, increasing $\Delta x$ {\it decreases} the value of the Laplacian term, while the local term is unchanged. For $\Delta x$ large enough the Laplacian cannot overcome the local term, the solution stops evolving, and the system artificially ``pins''. \section{Conclusion} \label{sec:conclusion} In summary, we have shown that both critical and off-critical polymer blends described by the Flory-Huggins-De Gennes free energy functional undergo spinodal decomposition via the Cahn-Hilliard equation without pinning of the domain growth as observed in some experiments. Even in the absence of thermal noise\cite{brown93}, the solution of the discretized equation of motion shows coarsening without evidence of pinning, regardless of the relative concentration of the blend components. We have also shown that previous solutions of the CH-FHDG equation that exhibited pinning were artifacts of an oversized mesh size used in the discretization and numerical integration of the equation of motion, and {\it not} a result of the concentration-dependence of the square gradient coefficient as previously suggested \cite{kotnis92}. This suggests the FHDG free energy functional alone, as written in Eq.~\ref{eq:fhdg} is not sufficient to describe the physics responsible for pinning in real blends. Evidently, a model able to describe the arrested growth observed in experiments must include additional physical ingredients. One should consider that experimental blends which exhibit pinning often contain components which are not simple homopolymers; such is the case with Hashimoto's random copolymer/homopolymer blends, as well as with Hasegawa's liquid-crystalline polymer/homopolymer blend, in which the liquid-crystalline component is anistropic at the quench temperatures at which pinning was observed. These blends may not be describable by the simple Flory-Huggins-De Gennes expression, and consequently we should not expect the CH-FHDG equation to mimic their behavior during spinodal decomposition. It is also important to note that a significant fraction of homopolymer blends never exhibit pinning, regardless of the relative composition. However, for those that do, it is possible that either the mobility, or free energy functional, or both, must be appropriately modified. The simplified expression for mobility (Eq.~\ref{eq:mob}) used in the CH-FHDG equation, which was originally derived only for the special case of a perfectly symmetric blend \cite{degennes80}, may neglect important contributions to mobility arising from the connectivity within the polymer chains. With respect to possible free energy modifications, it is well known that interfacial growth can be slowed or stopped by decreasing the interfacial tension. This can be achieved by, e.g. using surfactants in the case of small molecules\cite{surfactants}, diblock copolymers in the case of polymers\cite{copolymers}, or impurities in the case of alloys\cite{impurities}. If the experimental blends contain even a small number of impurities, or specific interaction regions that act as impurities \cite{ggscs94}, these impurities could arrest the demixing process and cause pinning. Finally, the FHDG free energy functional in Eq.~\ref{eq:fhdg} describes incompressible blends; real blends are in fact compressible. A coupling of concentration fluctuations and density fluctuations may be responsible for pinning in blends \cite{douglas}, in which case a reformulation of the free energy functional as well as the addition of a second order parameter field is necessary. We are extremely grateful to J. Douglas, B. Hammouda, P. Gallagher, C. Han, E. DiMarzio, G. McFadden, A. Coniglio, F. Corberi and J. Warren, and especially to M. Muthukumar, A. Chakrabarti and G. Brown, for useful discussions. We thank the Center for Computational Science at Boston University and the University of Maryland for generous use of their CM-5. CC would like to thank the Structure and Mechanics Group in the Polymers Division at NIST, and the NIST Center for Theoretical and Computational Materials Science, for their hospitality. \section{Appendix} The numerical solution of Eq.~\ref{eq:chfhdg2} is performed via iteration of the following map: \begin{eqnarray} \phi^{m+1}_{i,j,k} = \phi^m_{i,j,k} + \Delta \tau {\partial \phi^m_{i,j,k} \over \partial \tau}. \end{eqnarray} This map, given the value of the concentration field $\phi^m_{i,j,k}$ at time $m \Delta \tau$ at each of the $n^3$ sites of a simple cubic lattice with mesh size $\Delta x$, yields the value of $\phi^{m+1}_{i,j,k}$ at each site at time $(m+1) \Delta \tau$. Note that the mesh size is taken to be the same in all directions. The time derivative $\partial \phi^m_{i,j,k} / \partial \tau$ is given by the discretization of the left hand side of Eq.~\ref{eq:chfhdg2}, with spatial derivatives centrally discretized. This means the chemical potential $\mu^m_{i,j,k}$ that appears in square brackets in Eq.~\ref{eq:chfhdg2} is computed using: \begin{eqnarray} \left[\nabla \phi({\bf x},\tau)\right]^2 \to \left( {\phi^m_{i+1,j,k} - \phi^m_{i-1,j,k} \over 2\Delta x}\right)^2 + \left( {\phi^m_{i,j+1,k} - \phi^m_{i,j-1,k} \over 2\Delta x}\right)^2 + \left( {\phi^m_{i,j,k+1} - \phi^m_{i,j,k-1} \over 2\Delta x}\right)^2, \end{eqnarray} and \begin{eqnarray} \nabla^2 \phi({\bf x},\tau) \to {\phi^m_{i+1,j,k} - 2\phi^m_{i,j,k} + \phi^m_{i-1,j,k} \over (\Delta x)^2} + {\phi^m_{i,j+1,k} - 2\phi^m_{i,j,k} + \phi^m_{i,j-1,k} \over (\Delta x)^2} + \cr \cr {\phi^m_{i,j,k+1} - 2\phi^m_{i,j,k} + \phi^m_{i,j,k-1} \over (\Delta x)^2}. \end{eqnarray} The divergence of the product $\phi (1-\phi) \nabla \mu$ is evaluated as follows: \begin{eqnarray} \nabla \cdot \{ \phi (1-\phi) \nabla \mu \} \to \left( {X^m_{i+1,j,k} - X^m_{i-1,j,k} \over 2\Delta x} \right) + \left( {Y^m_{i,j+1,k} - Y^m_{i,j-1,k} \over 2\Delta x} \right) + \cr \cr \left( {Z^m_{i,j,k+1} - Z^m_{i,j,k-1} \over 2\Delta x} \right), \end{eqnarray} where \begin{eqnarray} X^m_{i,j,k} = \left( {\mu^m_{i+1,j,k} - \mu^m_{i-1,j,k} \over 2\Delta x} \right) \left( \phi^m_{i,j,k} (1 - \phi^m_{i,j,k}) \right), \end{eqnarray} and $Y^m_{i,j,k}$ and $Z^m_{i,j,k}$ are defined accordingly.
\section{Introduction} The basic measurement in deeply inelastic scattering (DIS) is a measurement of the cross section \proc{ep}{eH} in terms of the structure function $F_2$, where $H$ stands for any hadronic system. A wealth of information upon the partonic structure of the proton and its dynamics have been obtained from structure function measurements. Measurements of the properties of the hadronic final state $H$ provide complementary information which cannot be obtained from inclusive structure functions. In the simple quark parton model (QPM) of DIS, a quark is scattered out of the proton by the virtual boson emitted from the scattering lepton. QCD modifies this picture. Partons may be radiated before and after the boson-quark vertex, and the boson may also fuse with a gluon inside the proton by producing a quark-antiquark pair (\fref{qcdgraphs}). In fact, the parton which is probed by the boson may be the end point in a whole cascade of parton branchings. This parton shower materializes in the hadronic final state, allowing experimental access to the dynamics governing the cascade. \begin{figure}[htb] \centering \begin{picture}(1,1) \put(0.,30.){QPM} \end{picture} \begin{picture}(1,1) \put(120.,30.){BGF} \end{picture} \epsfig{file=qpm.eps,width=4cm, bbllx=50pt,bblly=483pt,bburx=522pt,bbury=771pt,clip=} \epsfig{file=bgf.eps,width=4cm, bbllx=68pt,bblly=435pt,bburx=289pt,bbury=661pt,clip=} \epsfig{file=qcdc.eps,width=8 cm} \begin{picture}(1,1) \put(-230.,30.){QCDC} \end{picture} \begin{picture}(1,1) \put(-110.,30.){QCDC} \end{picture} \caption{\em Diagrams for DIS in $O(\mbox{$\alpha_s~$}^0)$ (quark parton model - QPM) and in $O(\mbox{$\alpha_s~$}^1)$: boson-gluon fusion (BGF) and QCD Compton (QCDC) processes.} \label{qcdgraphs} \end{figure} HERA has opened a new kinematic domain to study QCD in DIS, and most contributions in this working group were concerned with HERA physics. In HERA electrons of $E_e \approx 27 \mbox{\rm ~GeV~}$ collide with protons of $E_p=820 \mbox{\rm ~GeV~}$, resulting in a centre of mass energy of $\sqrt{s}\approx 300 \mbox{\rm ~GeV~}$. The kinematic region covered with the present data is roughly $10^{-4}<x<10^{-1}$, $7 \mbox{${\rm ~GeV}^2~$}<Q^2<5000 \mbox{${\rm ~GeV}^2~$}$ and $40\mbox{\rm ~GeV~}<W<300\mbox{\rm ~GeV~}$, where \mbox{$W~$}\ is the invariant mass of the hadronic system, \mbox{$Q^2~$}\ the negative 4-momentum transfer squared, and \xb the Bjorken scaling variable, to be identified with the proton momentum fraction carried by the scattering parton. HERA offers the opportunity to study the evolution of physics quantities over a large kinematic range. The large phase space available for hard QCD radiation, which can be treated in perturbative QCD, leads to prominent jets observable in the final state. Another area of recent interest is the kinematic regime at small $x$ ($x\mathrel{\mathpalette\@versim<} 10^{-3}$) but sizeable \mbox{$Q^2~$}, not accessible at pre-HERA DIS experiments, where novel QCD dynamics are expected to play a r\^{o}le (e.g. \cite{levin}). The two HERA detectors ZEUS and H1 \cite{detectors} are large multipurpose, ``almost 4$\pi$'' detectors built around the beam line. Inner tracking detectors for charged particle detection are surrounded by a magnet and calorimetry. Both the scattered electron serving as a tag for DIS events and the hadronic final state are measured. Note that a substantial part of the hadronic final state, the proton remnant, leaves the detectors unobserved in the beam pipe. The region close to the proton beam direction is often referred to as the forward region. Apart from the laboratory frame, the hadronic centre of mass system (CMS) and the Breit frame are used in the analyses. The Breit frame is defined by the condition that the virtual photon does not transfer energy, only momentum. In the QPM picture the scattering quark would thus just reverse its momentum of magnitude $Q/2$. The CMS is defined as the centre of mass system of the incoming proton and the virtual boson, i.e. the CMS of the hadronic final state with invariant mass $W$. In both systems the hemisphere defined by the virtual photon direction is referred to as the current region, the other (containing the proton remnant) as the target region. The CMS current and target systems are back to back with momentum $W/2$ each. Longitudinal and transverse quantities are calculated w.r.t. the boson direction. With a longitudinal boost from the Breit frame into the CMS, particles formerly assigned to the target hemisphere may now end up in the current hemisphere. Monte Carlo (MC) models based upon QCD phenomenology are used to simulate the DIS process. The MEPS model (Matrix Element plus Parton Shower), an option of the LEPTO generator \cite{lepto}, incorporates the QCD matrix elements up to first order, with additional soft emissions generated by adding leading log parton showers. In the colour dipole model (CDM) \cite{dipole,ariadne} radiation stems from a chain of independently radiating dipoles formed by the colour charges. Both programs use the Lund string model \cite{string} for hadronizing the partonic final state. Deficiencies of the Herwig parton shower model \cite{herwig} have now been fixed by adding matrix element corrections \cite{seymour,webber}. This model implements an alternative (cluster) fragmentation scheme \cite{cluster}, allowing for valuable cross checks in the future. \section{Jet physics} The processes contributing to DIS up to first order in \mbox{$\alpha_s~$} are shown in figure \ref{qcdgraphs}. The QPM process results in a so-called ``1+1'' jet topology, while the QCDC and BGF processes give ``2+1'' jet events, where the ``+1'' refers to the unobserved remnant jet. {}From a measurement of the 2+1 jet rate at large \xb and \mbox{$Q^2~$}, where the parton densities are well known, \mbox{$\alpha_s~$} can be measured. At small \xb and \mbox{$Q^2~$}, one can determine the largely unknown gluon density from the rate of 2+1 jet events, which is then dominated by the BGF graph (assuming \mbox{$\alpha_s~$} to be known). Complications arise from the fact that the initial state contains strongly interacting particles, leading to the evolution of parton showers. Such effects need to be taken into account with the help of MC simulations. \subsection{The strong coupling constant \mbox{$\alpha_s~$}} Both H1 and ZEUS use the modified JADE algorithm \cite{jade} with resolution parameter $\mbox{$y_{\rm cut}~$}=0.02$ to define jets in the \mbox{$\alpha_s~$} analysis. A pseudoparticle is introduced to account for the unobserved remnant, and then all particles $i,j$ satisfying $m_{ij}^2<\mbox{$y_{\rm cut}~$} \cdot W^2$ are merged into jets. The chosen \mbox{$y_{\rm cut}~$} value is a compromise between statistical precision (small \mbox{$y_{\rm cut}~$}), and controllable higher order corrections (large \mbox{$y_{\rm cut}~$}) \cite{graudenz}. In the H1 analysis \cite{h1alphas} an angular cut $\mbox{$\theta_{\rm jet}~$} > 10^\circ $ (w.r.t. the proton direction) protects against parton showers close to the remnant. The obtained jet rates are corrected for detector effects, remaining parton shower contributions and hadronization with the MEPS model. In order to extract \mbox{$\alpha_s~$} from the measured jet rates, it is important to take next to leading order (NLO) corrections into account to reduce dependencies upon \mbox{$y_{\rm cut}~$} and the chosen renormalization and factorization scales \cite{graudenz}. Using PROJET \cite{projet} as NLO calculation, the measured jet rate then yields measurements of $\mbox{$\alpha_s~$}(Q^2)$ in the range $10 \mbox{${\rm ~GeV}^2~$}< \mbox{$Q^2~$} < 3000 \mbox{${\rm ~GeV}^2~$}$, which can be seen to run according to the QCD expectation \cite{h1alphas}. However, below $\mbox{$Q^2~$}=100 \mbox{${\rm ~GeV}^2~$}$, the corrections are very model dependent (MEPS vs. CDM). Therefore only data at $\mbox{$Q^2~$}>100 \mbox{${\rm ~GeV}^2~$}$ are used to extract $\mbox{$\alpha_s~$}(\mbox{$m_Z~$}^2)=0.123 \pm 0.018$ \cite{h1alphas}. For 2+1 jet events with $\mbox{$Q^2~$}>160\mbox{${\rm ~GeV}^2~$}$ and $x>0.01$, ZEUS has measured the jet distribution in the Lorentz invariant \mbox{$z_p~$} variable \cite{zjets}, which in the centre of mass frame of the virtual photon and the incoming parton is an angular variable $\mbox{$z_p~$}=\frac{1}{2}\cdot (1-\cos \hat{\theta}_{\rm jet})$. Here $\hat{\theta}_{\rm jet}$ is the angle of the jet w.r.t. the direction of the incoming parton. Perturbation theory in next to leading order (NLO) \cite{projet} is able to describe the jet angular distribution down to $\mbox{$z_p~$} \approx 0.1$. For $\mbox{$z_p~$}<0.1$ an excess of jets is observed. Both, the MEPS (LO matrix element + parton showers) and ME (pure LO matrix element) simulations are similar to the NLO calculation \cite{zjets}. The excess of jets at $\mbox{$z_p~$}<0.1$ is therefore unlikely to be cured by next to NLO calculations. For the \mbox{$\alpha_s~$} extraction, a cut $\mbox{$z_p~$}>0.1$ restricts the data to a region well described by NLO perturbation theory and QCD models \cite{grindhammer}. The preliminary \mbox{$\alpha_s~$} measurements \cite{grindhammer} for $100 \mbox{${\rm ~GeV}^2~$} \mathrel{\mathpalette\@versim<} \mbox{$Q^2~$} \mathrel{\mathpalette\@versim<} 3600\mbox{${\rm ~GeV}^2~$}$ demonstrate the potential of HERA to study the dependence of \mbox{$\alpha_s~$} upon the renormalization scale, and agree well with the QCD expectation (see \fref{as}). It is expected that already the analysis of the 1994 HERA data, once finalized, will yield a very competitive measurement of $\mbox{$\alpha_s~$}(\mbox{$m_Z~$}^2)$. \subsection{The gluon density in the proton} The 2+1 jet sample (defined with the cone algorithm in the CMS) in the range $10\mbox{${\rm ~GeV}^2~$}<\mbox{$Q^2~$}<100 \mbox{${\rm ~GeV}^2~$}$ is used to extract the gluon density \mbox{$g(x_g,Q^2)$~}, because there the BGF graph (\fref{qcdgraphs}) dominates (BGF:QCDC $\approx 4:1$ \cite{grindhammer,h1gx}). The momentum fraction \mbox{$x_g~$} which the gluon carries is calculated from the invariant mass$^2$ \mbox{$\hat{s}~$} of the hard subsystem forming the 2 jets via $\mbox{$x_g~$}=x(1+\hat{s}/\mbox{$Q^2~$})\approx \hat{s}/W^2$. Special cuts remove events affected by parton showers \cite{grindhammer,h1gx}. The MEPS model is used to unfold detector effects, the QCDC contribution, QPM background and remaining parton shower contributions. The MEPS model employs a cut-off for invariant parton-parton masses $m_{ij}^2>\mbox{$y_{\rm min}~$}\cdot W^2$ to regulate divergencies of its LO matrix element. In order to access \mbox{$x_g~$} as small as possible, \mbox{$\hat{s}~$} is chosen as small as experimental resolution allows, and as problems with the diverging LO matrix element can be avoided. It has to be ensured that the BGF events to be analyzed are actually generated by the model and do not fall below that cut-off \cite{grindhammer,h1gx}. The H1 analysis \cite{grindhammer} uses a fixed cut-off $\hat{s}>100 \mbox{${\rm ~GeV}^2~$}$ to define BGF events, and they parametrize the MEPS cut-off such as to follow the limit at which the order \mbox{$\alpha_s~$} contribution exceeds the total cross section within a margin of $\Delta \sqrt{\hat{s}} = 2 \mbox{\rm ~GeV~}$. ZEUS uses the standard \mbox{$y_{\rm min}~$} cut-off scheme in the MEPS model and defines BGF events via $\hat{s}>\mbox{$y_{\rm min}~$} \cdot W^2$. The parameter \mbox{$y_{\rm min}~$} is then varied between 0.0025 and 0.01 to study its influence on the result. The H1 and ZEUS results \cite{grindhammer} agree well with each other, but yield different size systematic errors (figure~\ref{xgluon}). The ZEUS errors receive large contributions from the \mbox{$y_{\rm min}~$} variation. The rise of the measured gluon density towards small \xb can be described by a LO gluon density \cite{grv} following the DGLAP (Dokshitzer-Gribov-Lipatov-Altarelli-Parisi) \cite{dglap} equations. The data are also consistent with the indirect determination of \mbox{$g(x_g,Q^2)$~} from the scaling violations of $F_2$ \cite{qcdfit}, providing a non-trivial test of QCD. \begin{figure}[t] \centering \epsfig{file=alphas.ps,width=7cm, bbllx=52pt,bblly=188pt,bburx=494pt,bbury=610pt,clip=} \caption{\em Preliminary $\mbox{$\alpha_s~$} (Q)$ measurements from ZEUS, compared to the QCD predictions corresponding to \mbox{$\Lambda_{\rm \bar{MS}}~$} =~100, 200 and 300~GeV.} \label{as} \end{figure} \begin{figure}[t] \centering \begin{picture}(1,1) \put(160.,55.){prelim.} \end{picture} \epsfig{file=gx.ps,width=7cm, bbllx=25pt,bblly=291pt,bburx=520pt,bbury=631pt,clip=} \caption{\em The gluon density in the proton, determined in leading order (LO) from the rate of 2+1 jet events. Shown are data from H1 and ZEUS, compared to the LO GRV \protect\cite{grv} gluon density parametrization.} \label{xgluon} \end{figure} \subsection{Open Points} Lack of understanding of parton showers close to the remnant (model dependent corrections, failure of NLO calculations) currently prevents the \mbox{$\alpha_s~$} analysis to make full use of the large statistics data at $\mbox{$Q^2~$} < 100 \mbox{${\rm ~GeV}^2~$}$. Though increasing HERA luminosity will allow the \mbox{$\alpha_s~$} analysis to be restricted to higher \mbox{$Q^2~$} to reduce uncertainties, the understanding of the forward region remains a challenge. So far the \mbox{$\alpha_s~$} measurements rely solely upon the JADE algorithm, being the only algorithm for which NLO jet cross sections are calculated \cite{nlo}. NLO calculations for other algorithms, such as the cone \cite{cone} or the theoretically preferred $k_T$ \cite{kt} algorithm are desirable. Such a program, which would also be able to calculate event shape variables like energy-energy correlations, Thrust, etc., is being worked upon by D. Graudenz, but results cannot be expected in a short term. Theoretical uncertainties could also be reduced by resumming higher order corrections. The validity of corrections from hadronic to partonic final states, defined either in LO or NLO, need to be checked with models based upon different parton shower and hadronization schemes. Unfortunately, a MC generator incorporating the QCD matrix elements beyond LO is missing. The gluon density has so far been determined in LO. A method allowing a measurement in NLO is presently under study \cite{graudenz}. How can \mbox{$\alpha_s~$} be determined consistently, considering it is input for the evolution of parton densities which are used in the analysis \cite{vogt}? \section{Novel QCD dynamics} The observed strong rise of the structure function $F_2$ towards small $x$ \cite{f2} has caused much debate on whether the QCD evolution of the parton densities can still be described by the conventional DGLAP \cite{dglap} equations, or whether the HERA data extend into a new regime at small $x$ where the dynamics is governed by the BFKL (Balitsky-Fadin-Kuraev-Lipatov) \cite{bfkl} equation. It would be extremely interesting to test QCD in such a new regime. While the rise is consistent with the expectation from BFKL dynamics, it can however also be described by a DGLAP evolution \cite{akms}. At lowest order the BFKL and DGLAP equations resum the leading logarithmic $(\mbox{$\alpha_s~$} \ln 1/x)^n$ or $(\mbox{$\alpha_s~$} \ln (\mbox{$Q^2~$}/ Q_0^2))^n$ contributions respectively. In this approximation the leading diagrams are of the ladder type (\fref{cascade}). The leading log DGLAP ansatz corresponds to a strong ordering of the transverse momenta \mbox{$k_T~$} (w.r.t. the proton beam) in the parton cascade ($Q_0^2 \ll \mbox{$k_T~$}_1^2 \ll ... \mbox{$k_T~$}_j^2 \ll ... \mbox{$Q^2~$}$), while there is no such ordering in the BFKL ansatz ($\mbox{$k_T~$}_j^2 \approx \mbox{$k_T~$}_{j+1}^2$) \cite{ordering}. Measurements on the hadronic final state emerging from the cascade therefore offer another handle to search for signatures of the BFKL behaviour. They are compared to analytical calculations as well as to the QCD models MEPS and CDM. The CDM description of gluon emission is similar to that of the BFKL evolution, because the gluons emitted by the dipoles do not obey strong ordering in \mbox{$k_T~$}~\cite{bfklcdm}. The MEPS model with its leading log parton shower is based upon DGLAP dynamics, and the emitted partons are thus ordered in \mbox{$k_T~$}. \begin{figure}[t] \centering \vspace{-0.2cm} \epsfig{file=cascade.eps,width=2cm} \caption{\em Parton evolution in the ladder approximation. The selection of forward jets in DIS events is illustrated.} \label{cascade} \end{figure} \subsection{Transverse Energy Production} As a consequence of the strong \mbox{$k_T~$} ordering the DGLAP evolution is expected to produce less transverse energy \mbox{$E_T~$} in a region between the current region and the proton remnant than the BFKL evolution \cite{durham}. H1 and ZEUS have measured the flow of transverse energy in the laboratory frame as a function of pseudorapidity $\eta = - \ln \tan (\theta/2)$, where $\theta$ is the angle of the energy deposition w.r.t the proton beam axis \cite{h1flow2,haas,h1flow3}. The measurements are made for varying ranges in \xb ($2 \mbox{$\cdot 10^{-4}$} < \av{\xb} < 5 \mbox{$\cdot 10^{-3}$}$) and \mbox{$Q^2~$} ($7\mbox{${\rm ~GeV}^2~$} < \av{\mbox{$Q^2~$}} < 30 \mbox{${\rm ~GeV}^2~$}$ and agree well between the experiments \cite{haas}. The \mbox{$E_T~$} flows for large \xb and \mbox{$Q^2~$} are reasonably well described by MEPS and CDM. For smaller \xb and \mbox{$Q^2~$} both models predict a more pronounced enhancement in the current fragmentation region than is seen in the data. Between the current system and the proton remnant (the lab. forward region), the data are reasonably well described by the CDM, while the MEPS model produces too little \mbox{$E_T~$} \cite{h1flow2,haas}. This intermediate region is expanded in \fref{et93}, because there perturbative calculations, based either on DGLAP or on BFKL dynamics, are available \cite{durham}. The BFKL calculation comes out close to the data, while the DGLAP calculation predicts much less \mbox{$E_T~$}. However, the non-perturbative hadronization phase is missing in these calculations. \ffig{et93.eps}{60mm} {\em Transverse energy flow in the forward region at H1 \protect\cite{h1flow2} and ZEUS \protect\cite{haas} for $x<10^{-3}$. The proton direction is to the right. The calorimeter acceptances end at $\eta$ around 3.5. The data are compared to the CDM (here labelled CDMBGF) and MEPS models and to partonic calculations based upon the DGLAP and BFKL equations \protect\cite{durham}.} {et93} H1 has determined the average \mbox{$E_T~$}, measured centrally in the CMS as a function of \xb and \mbox{$Q^2~$} (\fref{etx}). They find an increase of \av{\mbox{$E_T~$}} with decreasing \xb, which is a characteristic BFKL prediction \cite{durham}. The data are in agreement with the BFKL calculation \cite{sutton}, if one assumes an \mbox{$E_T~$} contribution from hadronization of about 0.4 GeV per unit rapidity (independent of \xb). That estimate is taken from the CDM, which agrees with the BFKL calculation at the parton level. \ffig{etx.eps}{60mm} {\em Transverse energy \av{\mbox{$E_T~$}} per unit of pseudorapidity $\eta^\ast$ as a function of \xb for three different values of \mbox{$Q^2~$}, measured centrally at $\eta^\ast = 0$ in the CMS (corresponding to the lab. forward region). The data are compared to the CDM and MEPS models including hadronization, and to the BFKL calculation (no hadronization).} {etx} The apparent failure of the MEPS model has caused many questions about its ingredients: the way the parton shower is ``matched'' to the matrix element, the colour connection between the current and the remnant system and its effect upon hadronization, and the remnant fragmentation itself which is little tested. It seems that re-arranging colour configurations can produce enough \mbox{$E_T~$} through hadronization to compensate the \mbox{$E_T~$} deficit in the DGLAP cascade of the MEPS model \cite{ingelman}. A MEPS version thus modified should be available soon for detailed testing. The flexibility in the hadronization modelling presently precludes unambiguous tests of the DGLAP evolution through \mbox{$E_T~$} measurements. For the same reasons the intriguing success of the CDM without \mbox{$k_T~$} ordering may be fortuitous. A MC model invoking explicitly the BFKL evolution, currently being developed by K. Golec-Biernat et al., would help interpreting the data. In any case, the \mbox{$E_T~$} data provide important input for QCD phenomenology. \subsection{Forward Jets} At present strong conclusions upon the validity of the BFKL or DGLAP parton evolutions at small $x$ from the \mbox{$E_T~$} measurements are hampered by the uncertainties about hadronization. Jet production should be less affected by hadronization. A signature for BFKL dynamics proposed by \cite{mueller} is the production of ``forward jets'' with $\mbox{$x_{\rm jet}$}=\mbox{$E_{\rm jet}$}/E_p$, the ratio of jet energy and proton beam energy, as large as possible, and with transverse momentum \mbox{$k_{T {\rm jet}}$} ~close to \mbox{$Q~$} in order to reduce the phase space for the \mbox{$k_T~$} ordered DGLAP evolution (see \fref{cascade}). An enhanced rate of events with such jets is thus expected in the BFKL scheme \cite{mueller,dhotref}. The experimental difficulty is to detect these ``forward'' jets which are close to the beam hole in proton direction. The rate of forward jets measured by H1 \cite{deroeck,h1flow3} (\fref{fwdjets}) is larger at low \xb than at high $x$. This is expected from BFKL calculations, in contrast to calculations without BFKL ladder \cite{dhotref,delduca}. The behaviour of the data is better represented by the CDM than by the MEPS model. However, neither of them describe the energy spectrum of the observed jets correctly, and the model predictions for the jet rates are thus cut dependent \cite{deroeck,h1flow3}. The analysis of a larger statistics sample should allow more firm conclusions. \begin{figure}[t] \centering \begin{picture}(1,1) \put(40.,40.){preliminary} \end{picture} \epsfig{file=fwdjets.eps,width=5cm} \caption{\em The rate of forward jets (selected with $\mbox{$x_{\rm jet}$}>0.025$, $0.5 < \mbox{$k_{T {\rm jet}}$}^2/\mbox{$Q^2~$} <4$ and $\mbox{$k_{T {\rm jet}}$}>5 \mbox{\rm ~GeV~}$) in the kinematic range $2 \mbox{$\cdot 10^{-4}$} < \xb < 2 \mbox{$\cdot 10^{-3}$}$ and $\mbox{$Q^2~$} \approx 20 \mbox{${\rm ~GeV}^2~$}$. The measurement is compared to the CDM and MEPS models.} \label{fwdjets} \end{figure} ZEUS has measured an inclusive jet cross section ${\rm d}\sigma / {\rm d} \mbox{$E_{T {\rm jet}}$}$ in the Breit frame \cite{deroeck}. Many more jets are found in the target region with a harder \mbox{$E_{T {\rm jet}}$}~ spectrum than in the current region, reflecting the differences in phase space in the two regions. The current region data are reasonably well described by the CDM and MEPS models. In the target region however there is a substantial excess of jets over the model predictions, which can be linked with an excess of $2+1$ jet events \cite{deroeck}. In the laboratory frame this excess is located in the forward region at angles $\theta_{\rm jet}<20^\circ$ (\fref{thjet}). The data on jet production in the forward region (lab. frame), or the target region (Breit frame) certainly pose a challenge to theory. So far cross sections are calculated \cite{dhotref,delduca} only for partons, while experiments measure hadron jets. This gap has to be bridged from both sides to allow a strictly valid comparison. \begin{figure}[t] \centering \vspace{-3.8cm} \begin{picture}(1,1) \put(80.,90.){ZEUS preliminary} \end{picture} \epsfig{file=thnew.ps,width=8cm, bbllx=22pt,bblly=165pt,bburx=522pt,bbury=645pt,clip=} \caption{\em The laboratory angular distribution of jets detected either in the Breit current or target hemisphere.} \label{thjet} \end{figure} \subsection{Jet correlations} Apart from calculations of forward jet rates, Del Duca \cite{delduca} discussed angular correlations for forward jets. If a BFKL ladder is inserted in between the electron-photon vertex and the forward jet, the angular correlation between the forward jet and the electron imposed by momentum conservation is relaxed. Such angular decorrelation could be another footprint of BFKL dynamics. The fact that 4\% of the H1 forward jet events contain a second forward jet \cite{deroeck} opens another route of investigation, namely correlations between such jets. If these jets can be identified with gluons emitted from the ladder, it would be possible to check the parton ordering directly. \subsection{Dipole emission} An interesting ansatz to calculate final state observables was presented by R. Peschanski \cite{peschanski}. The starting point is onium-onium scattering \cite{onium} with onium wave functions which can be derived from QCD. Such a reaction is analogous to an interaction of the current system with the remnant system in DIS. Radiation is treated in the dipole picture, leading to a copious production of dipoles in the central rapidity region of the interaction. Once such an ansatz yields quantitative predictions, it could be tested in DIS, e.g. with \mbox{$E_T~$} flow measurements. Bo Andersson \cite{andersson} discussed DIS final states in terms of a chain of radiating colour dipoles, and its connection with the Ciafaloni-Catani-Fiorini-Marchesini ansatz \cite{ccfm}. In principle this model could provide a complete picture of the hadronic final state in DIS. The implementation in the Ariadne \cite{ariadne} MC generator is in progress to allow detailed predictions. \subsection{QCD Instantons} The standard model contains processes which cannot be described by perturbation theory, and which violate classical conservation laws like baryon and lepton number in the case of the electroweak sector and chirality for the strong interaction \cite{thooft}. Such anomalous processes are induced by instantons \cite{belavin}. At HERA, QCD instantons may lead to observable effects in the hadronic final state in DIS \cite{ringwald,schrempp}, which were discussed by F. Schrempp. The instanton should decay isotropically into a high multiplicity state of gluons and all quark flavours simultaneously which are kinematically allowed. A MC program to simulate instanton events has become available \cite{gibbs}. Due to the isotropic decay, one expects a densely populated region in rapidity, other than the current jet, which is isotropic in azimuth. The presence of strangeness and charm could provide an additional signature. However remote the a priori chances to see such signals may appear, here is a chance for a major discovery at HERA! \section{Charged Particle Spectra} The H1 and ZEUS measurements of inclusive charged particle spectra \cite{pavel} are performed either in the Breit frame or in the CMS. In the Breit frame in- and outgoing quark have equal but opposite sign momenta $Q/2$ (QPM picture), and in \mbox{$e^+e^-$} annihilation the outgoing quark and antiquark have equal but opposite momenta $\sqrt{s}/2=Q/2$. Due to this similarity it is interesting to compare particle spectra in the Breit current hemisphere in DIS with \mbox{$e^+e^-$} data. DIS experiments have the advantage over \mbox{$e^+e^-$} experiments that they cover a large span in \mbox{$Q~$}, presently from 3 \mbox{\rm ~GeV~} to 50 \mbox{\rm ~GeV~}, in a single experiment. The current mean charged multiplicity at HERA rises $\sim \ln Q$ within errors, and agrees with \mbox{$e^+e^-$} data (divided by 2) where they overlap \cite{breit,pavel}. Colour coherence should lead to a suppression of soft gluon emission. The HERA data \cite{breit,pavel} on the scaled charged particle momentum distribution $\ln 1/x_p$ with $x_p=2\cdot p/Q$ exhibit the expected hump backed plateau \cite{basics}, the evolution of which with \mbox{$Q~$} is in agreement with the assumption of colour coherence. However, like in \mbox{$e^+e^-$} annihilation, this behaviour can also be mimicked through the Lund string fragmentation \cite{pavel}. The scaled momentum spectrum of \xf in the CMS, where the particle longitudinal momenta $p_z$ are divided by the maximal possible momentum, $\xf=2 \cdot p_z/W$, are shown in \fref{xf} for the current region (the target region is not observed). Comparing HERA data at $W \approx 120 \mbox{\rm ~GeV~}$ \cite{h1flow2,pavel} with fixed target data at \mbox{$W~$} = 14 and 18 \mbox{\rm ~GeV~} \cite{emc,e665}, significant scaling violations are observed, in agreement with QCD expectations: the large value of \mbox{$W~$} at HERA results in a large phase space for QCD radiation, softening the \xf spectrum w.r.t. data at lower $W$. It can be expected that such data will be used to extract \mbox{$\alpha_s~$} in the future. \ffig{xf.eps}{80mm} {\em The \xf spectra measured at HERA compared with the QPM (dotted line) without QCD radiation, the MEPS model (full line), and with fixed target DIS data at lower $W$.} {xf} The effect of QCD radiation is clearly seen in the ``seagull plot'' (\fref{sg}), where the mean transverse momenta \mbox{$p_T^{2}~$} squared of the particles is plotted as a function of \xf. As a consequence of increased QCD radiation, much larger \mbox{$p_T^{2}~$} are observed at HERA \cite{h1flow2,pavel} than at EMC \cite{emc} at smaller $W$, again in agreement with QCD expectation. ZEUS has also compared DIS events with and without a large rapidity gap \cite{gap} in this respect \cite{pavel}. Much smaller $p_T^2$ than in normal DIS events are observed in events with a large rapidity gap, thought to stem from diffractive processes and accounting for approximately 10\% of the total sample \cite{gap}. This indicates that the scale governing radiation is much smaller than \mbox{$W~$} for rapidity gap events. \section{Conclusion} Two complementary approaches to the HERA data can be distinguished. In one approach, one tries to identify a region which is ``well understood'', meaning that the observation agrees with the theory and the models. Under this condition, the data can be interpreted in the framework of the theory, and physical quantities which are defined within the theory can be extracted. The measurements of \mbox{$\alpha_s~$} and \mbox{$g(x_g,Q^2)$~} fall into this category. However, we have also seen data which are not yet understood theoretically, namely hadron and jet production in the forward region. Such data currently pose a challenge to the theory, and experimentalists should make every effort to provide theory with solid data to work with. \begin{center} {\large\bf Acknowledgements} \end{center} I would like to thank my fellow conveners, A. Doyle and G. Ingelman, for the pleasant cooperation, the organizers of the workshop for their efficient support and the participants of the session for their contributions and inspiring discussions in the working group. \begin{figure}[t] \centering \epsfig{file=sg_new.eps,width=68mm} \caption{\em The seagull plot. Shown are the mean transverse momenta squared $\av{p_T^2}$ as a function of \xf in the CMS for HERA data with and without a rapidity gap (LRG/NRG) compared to the QPM prediction (dotted line) and the MEPS model (full line), and to EMC data at lower $W$.} \label{sg} \end{figure} \Bibliography{100} \bibitem{levin} J. Bartels and J. Feltesse, Proc. of the Workshop on Physics at HERA, Hamburg 1991, eds. W. Buchm\"uller and G. Ingelman, vol. 1, p. 131;\\ E.M. Levin, Proc. QCD -- 20 Years Later, Aachen 1992, eds. P.M. Zerwas, H.A. Kastrup, vol. 1, p. 310. \bibitem{detectors} H1 Collab., I. Abt et al., DESY 93-103 (1993);\\ ZEUS Collab., M. Derrick et al., Phys. Lett. B293 (1992) 465. \bibitem{lepto} G. Ingelman, Proc. of the Workshop on Physics at HERA, Hamburg 1991, eds. W. Buchm\"uller and G. Ingelman, vol. 3, p. 1366. \bibitem{dipole} G. Gustafson, Ulf Petterson, Nucl. Phys. B306 (1988); \\ G. Gustafson, Phys. Lett. B175 (1986) 453; \\ B. Andersson, G. Gustafson, L. L\"onnblad, Ulf Petterson, Z. Phys. C43 (1989) 625. \bibitem{ariadne} L. L\"onnblad, Comp. Phys. Comm. 71 (1992) 15. \bibitem{string} T. Sj\"ostrand, Comp. Phys. Comm. 39 (1986) 347; \\ T. Sj\"ostrand and M. Bengtsson, Comp. Phys. Comm. 43 (1987) 367; T. Sj\"ostrand, CERN-TH-6488-92 (1992). \bibitem{herwig} G. Marchesini, B.R. Webber, G. Abbiendi, I.G. Knowles, M.H. Seymour and L. Stanco, Comp. Phys. Comm. 67 (1992) 465. \bibitem{seymour} M. Seymour, Lund preprint LU-TP-94-12 (1994). \bibitem{webber} B. Webber, these proceedings. \bibitem{cluster} B.R. Webber, Nucl. Phys. B238 (1984) 492. \bibitem{jade} JADE Collab., W. Bartel et al., Z. Phys. C33 (1986) 23. \bibitem{graudenz} D. Graudenz, these proceedings. \bibitem{h1alphas} H1 Collab., T.~Ahmed et al., Phys. Lett. B346 (1995) 415. \bibitem{projet} D. Graudenz, Projet 4.13 manual, CERN-TH 7420/94. \bibitem{zjets} ZEUS Collab., M. Derrick et al., DESY 95-016. \bibitem{grindhammer} G. Grindhammer, these proceedings. \bibitem{h1gx} H1 Collab., S. Aid et al., DESY 95-086 (1995). \bibitem{grv} M. Gl\"uck, E. Reya, A. Vogt, U. Dortmund preprint DO-TH-94-24. \bibitem{qcdfit} ZEUS Collab., M. Derrick et al., Phys. Lett. B345 (1995) 576;\\ H1 Collab., S. Aid et al., DESY 95-081 (1995). \bibitem{f2} ZEUS Collab., M. Derrick et al., Z. Phys. C65 (1995) 379; H1 Collab., T. Ahmed et al., Nucl. Phys. B439 (1995) 471. \bibitem{dglap} Yu. L. Dokshitzer, Sov. Phys. JETP 46 (1977) 641; \\ V.N. Gribov and L.N. Lipatov, Sov. J. Nucl. Phys. 15 (1972) 438 and 675; \\ G. Altarelli and G. Parisi, Nucl. Phys. 126 (1977) 297. \bibitem{nlo} D. Graudenz, Phys. Lett. B256 (1991) 518; Phys. Rev. D49 (1994) 3291; \\ T. Brodkorb, J.G. K\"orner, Z. Phys. C54 (1992) 519; \\ T. Brodkorb, E. Mirkes, U. Wisconsin preprint MAD/PH/820 (1994). \bibitem{cone} B. Webber, J. Phys. G19 (1993) 1567. \bibitem{kt} S. Catani, Y.L. Dokshitzer, B. Webber, Phys. Lett. B285 (1992) 291. \bibitem{vogt} A. Vogt, DESY 95-068. \bibitem{bfkl} E.A. Kuraev, L.N. Lipatov and V.S. Fadin, Sov. Phys. JETP 45 (1972) 199; \\ Y.Y. Balitsky and L.N. Lipatov, Sov. J. Nucl. Phys. 28 (1978) 282. \bibitem{akms} A.J. Askew, J. Kwieci\'{n}ski, A.D. Martin and P.J. Sutton, Phys. Lett. B325 (1994) 212. \bibitem{ordering} J. Bartels, H. Lotter, Phys. Lett. B309 (1993) 400; \\ A. Mueller, Columbia preprint CU-TP-658 (1994). \bibitem{bfklcdm} A. H. Mueller, Nucl. Phys. B415 (1994) 373;\\ L. L\"onnblad, Z. Phys. C65 (1995) 285 and CERN-TH/95-95. \bibitem{durham} J. Kwieci\'{n}ski, A.D. Martin, P.J. Sutton and K. Golec-Biernat, Phys. Rev. D50 (1994) 217. \\ K. Golec-Biernat, J. Kwieci\'{n}ski, A.D. Martin and P.J. Sutton, Phys. Lett. B335 (1994) 220. \bibitem{h1flow2} H1 Collab., I. Abt et al., Z. Phys. C63 (1994) 377. \bibitem{haas} T. Haas, these proceedings. \bibitem{h1flow3} H1 Collab., S. Aid et al., DESY-95-108. \bibitem{sutton} calculation by P. Sutton on the basis of \cite{durham}. \bibitem{ingelman} G. Ingelman, these proceedings. \bibitem{mueller} A.H. Mueller, Nucl. Phys. B (Proc. Suppl.) 18C (1990) 125; J. Phys. G17 (1991) 1443. \bibitem{dhotref} J. Kwieci\'{n}ski, A.D. Martin, P.J. Sutton, Phys. Rev. D46 (1992) 921. \bibitem{deroeck} A. DeRoeck, these proceedings. \bibitem{delduca} V. Del Duca, these proceedings. \bibitem{andersson} B. Andersson, these proceedings \bibitem{ccfm} M. Ciafaloni, Nucl. Phys. B296) (1988) 49; \\ S. Catani, F. Fiorani and G. Marchesini, Phys. Lett. B234 (1990) 339; Nucl. Phys. B336 (1990) 18. \bibitem{peschanski} R. Peschanski, these proceedings. \bibitem{onium} A.H. Mueller, Nucl. Phys. B415 (1994) 373; ibid. B437 (1995) 107. \\ A.H. Mueller and B. Patel, Nucl. Phys. B425 (1994) 471.\\ A. Bialas and R. Peschanski, Saclay-Orsay preprint T95/032, LPTHE-95/29. \bibitem{thooft} G. 't Hooft, Phys. Rev. Lett. 37 (1976) 8; Phys. Rev. D14 (1976) 3432. \bibitem{belavin} A. Belavin, A. Polyakov, A. Schwarz and Yu. Tyupkin, Phys. Lett. B59 (1975) 85. \bibitem{ringwald} A. Ringwald, Nucl. Phys. B330 (1990) 1; \\ O. Espinosa, Nucl. Phys. B343 (1990) 310. \bibitem{schrempp} A. Ringwald and F. Schrempp, DESY 94-197. \bibitem{gibbs} M. Gibbs, A. Ringwald and F. Schrempp, work presented by F. Schrempp at this workshop. \bibitem{pavel} N. Pavel, these proceedings. \bibitem{breit} ZEUS Collab., M. Derrick et al., DESY 95-007;\\ H1 Collab., I. Abt et al., DESY 95-072. \bibitem{basics} Y. Dokshitzer, V. Khoze, A. Mueller and S. Troyan, ``Basics of Perturbative QCD'', Gif-sur-Yvette, France (1991). \bibitem{e665} E665 Collab., M.R. Adams et al., Phys. Rev. D50 (1994) 1836. \bibitem{emc} EMC Collab., J. Ashman et al., Z. Phys. C52 (1991) 361. \bibitem{gap} ZEUS Collab., M. Derrick et al., Phys. Lett. B315 (1993) 481;\\ H1 Collab., T. Ahmed et al., Nucl.Phys. B429 (1994) 477. \end{thebibliography} \end{document}
\section{Introduction} Chiral perturbation theory is an effective theory which obeys the symmetries of QCD and contains a number of parameters which must be determined experimentally. If the theory reflects nature, then the parameters should be universal. This can be tested through one-loop SU(3) breaking calculations for the decays of the octet and decuplet baryons. Uncalculable terms may yield corrections of up to 30 percent to these predictions, but if the variance in the comparison to experiment goes beyond this, the validity of the chiral expansion is questioned for that process, and the reliability of estimating unmeasured processes is open. The two-body weak $\Delta s=1$ decays of hyperons are a natural place to investigate the validity of chiral perturbation theory. The experimental observables have been well measured, and calculations including leading logarithmic corrections (which appear through one-loop SU(3) breaking diagrams) have been performed\cite{Wise,EJ}. A comparison of these results, however, showed that the parameters which fit the S-wave decays was inadequate for describing the P-wave decays. This caused concern about the legitimacy of the chiral Lagrangian expansion, at least for these processes\cite{Wise,Georgi}. In this paper, previously omitted diagrams have been included in the calculation of P-wave amplitudes for nonleptonic hyperon decay. A correlated fit to the three weak parameters is performed and compared to previous results. The fit to the data improves markedly when one of the strong parameters, which is not well constrained at present, is also allowed to vary. \section{The Chiral Lagrangian for Nonleptonic Decays.} Heavy Baryon Chiral Perturbation Theory (HBChPT), which is used to make predictions for hadronic processes at momentum transfers much less than one GeV, is introduced and well described in Ref.\cite{mj}. The weak interaction portion of the Lagrangian needed for $\Delta s=1$ hyperon decays, which transforms under ${\rm SU(3)_L}\otimes {\rm SU(3)_R}$ as an $(8_L,1_R)$, is outlined in Ref.\cite{Wise,EJ}. The Lagrangian $${\cal L} = {\cal L}_{strong} + {\cal L}_{weak}$$ contains the particles which are dynamic in the energy regime relevant for hyperon decay. This includes the lowest mass octet and decuplet of baryons, and the octet of pseudo-Goldstone bosons. \begin{eqnarray}\label{strong} {\cal L}_{strong} &=& i \ {\rm Tr}\ \bar B_v\ \left(v\cdot {\cal D} \right)B_v + 2\ D\ {\rm Tr}\ \bar B_v\ S_v^\mu\ \{ A_\mu, B_v \} + 2\ F\ {\rm Tr}\ \bar B_v\ S_v^\mu\ [A_\mu, B_v] \nonumber \\ &&-\ i\ \bar T_v^{\mu}\ (v \cdot {\cal D}) \ T_{v \mu} + \Delta m\ \bar T_v^{\mu}\ T_{v \mu} + {\cal C}\ \left(\bar T_v^{\mu}\ A_{\mu}\ B_v + \bar B_v\ A_{\mu}\ T_v^{\mu}\right){\phantom {f^2 \over 4}} \nonumber\\ && +\ 2\ {\cal H}\ \bar T_v^{\mu}\ S_{v \nu}\ A^{\nu}\ T_{v \mu} + {f^2 \over 8}\ {\rm Tr}\ \partial_\mu \Sigma \partial^\mu \Sigma^\dagger + \mu\ {\rm Tr} \left( m_q\Sigma + m_q^\dagger\Sigma^\dagger \right) \ +\ \cdots \ \ \ \ , \end{eqnarray} where $f \sim 93$ MeV is the meson decay constant, the light quark mass matrix $m_q={\rm diag}\{m_u,m_d,m_s\}$, and ${\cal D_\mu}= \partial_\mu+[V_\mu, \; ]$ is the covariant chiral derivative. The subscript $v$ on the baryon fields makes explicit that, in HBChPT, velocity is a good quantum number and labels the field for this portion of the Lagrangian. The actual full Lagrangian is a sum over all such velocities on terms like that above. The $B_v$ are the octet of baryons, and the $T_v^\mu$ are the decuplet of baryons (the $\mu$ index is the Lorentz superscript for this Rarita-Swinger field). The vector and axial vector chiral currents used are defined by \begin{eqnarray} V_\mu&=&{1 \over 2} (\xi\partial_\mu\xi^\dagger + \xi^\dagger\partial_\mu\xi) \nonumber \\ A_\mu&=&{i \over 2} (\xi\partial_\mu\xi^\dagger - \xi^\dagger\partial_\mu\xi) \ \ \ . \end{eqnarray} Higher dimension operators, which contain more derivatives or insertions of the light quark mass matrix, are not needed in \eqn{strong} to the order we are working. The octet of pseudo-Goldstone bosons, $M$, appears through \begin{eqnarray} \Sigma = \xi^2= {\rm exp}\left( {2 i M\over f} \right) \ \ \ . \end{eqnarray} The strong couplings constants $F, D, {\cal C}$, and ${\cal H}$ have been obtained by comparing one-loop computations of axial matrix elements between octet baryons to semileptonic baryon decay measurements\cite{mj}. The constants ${\cal C}$ and ${\cal H}$ are further constrained through the one-loop computation of the strong decays of decuplet baryons \cite{bss}. This yields, \begin{eqnarray}\label{strongFDCH} D = 0.6\pm 0.1& \hskip 2cm & F = 0.4\pm 0.1\ \ \nonumber \\ 1.1 < |{\cal C}| < 1.8& \hskip 2cm &-2.8 < {\cal H} < -1.6 \ \ \ \ , \end{eqnarray} \noindent Note that the sign of ${\cal C}$ remains a convention. The errors do not include theoretical errors. Assuming octet dominance (the $\Delta I={1 \over 2}$ rule), the $\Delta s = 1$ weak Lagrangian is \begin{eqnarray}\label{weak} {\cal L}_{weak} &=& G_F\ m_\pi^2\ \sqrt{2} f_\pi\ h_D\ {\rm Tr}\ {\overline B}_v\ \lbrace \xi^\dagger h\xi \, , B_v \rbrace \; + \; G_F\ m_\pi^2\ \sqrt{2} f_\pi\ h_F\ {\rm Tr}\ {\overline B}_v\ {[\xi^\dagger h\xi \, , B_v ]} \; \nonumber \\ && + G_F\ m_\pi^2\ \sqrt{2} f_\pi\ h_C\ {\overline T}^\mu_v\ (\xi^\dagger h\xi)\ T_{v \mu} \; + \; G_F\ m_\pi^2\ h_\pi\ { f_\pi^2 \over 4}\ {\rm Tr} \left( h \, \partial_\mu \Sigma \partial^\mu \Sigma^\dagger \right) \ + \ \cdots\ , \end{eqnarray} where \begin{eqnarray} h = \left(\matrix{0&0&0\cr 0&0&1\cr 0&0&0}\right) \ \ \ , \end{eqnarray} \noindent picks out just the $\Delta s=1$ piece needed for hyperon decays. The constants $f_\pi$, $h_D, h_F, h_\pi$ and $h_C$ are then fit to reproduce experimental data. Predictive power is obtained because there are many more observables than parameters. The pion decay constant $f_\pi \sim 93$ MeV. Factors of $ G_Fm_\pi^2 \sqrt{2} f_\pi $ are inserted in \eqn{weak}\ so that the constants $h_D$, $h_F$, and $h_C$ are dimensionless. Nonleptonic kaon decays suggest that the weak meson coupling $h_\pi = 1.4$. \section{Hyperon Decay Amplitudes} In this section, the formulae for the S-wave and P-wave amplitudes for $\Delta S=1$ nonleptonic hyperon decay are discussed. The S-wave amplitudes were calculated previously\cite{EJ}. The portion of the P-wave amplitudes coming from the diagrams in Figure \ref{P1} were also calculated in \cite{EJ}. The pieces which are new and the subject of this work affect the P-wave amplitudes and arise from the diagrams in Figure \ref{P2}. The total amplitude for a decay of an initial octet baryon to a final octet baryon, $B_i \rightarrow B_f \pi$ is given by \begin{eqnarray}\label{spamp} {\cal A} = i\ G_F\ m_\pi^2\ \sqrt{2}\ f_\pi\ \overline{u}_{B_f} \ \left[ {\cal A}^{(S)} + 2 k \cdot S_v {\cal A}^{(P)} \right] u_{B_i} \ \ \ , \end{eqnarray} where $k$ is the outgoing momentum of the pion and $S_v$ is the spin operator for the baryons. The amplitudes ${\cal A}^{(S)}$ and ${\cal A}^{(P)}$ are the S-wave and P-wave amplitudes. Of all physically possible decays within the octet of baryons, only four are independent after isospin symmetry has been imposed. In keeping with Refs. \cite{Wise,EJ}, we will continue to choose those four to be $\Sigma^+ \rightarrow n \pi^+$, $\Sigma^- \rightarrow n \pi^-$, $\Lambda \rightarrow p \pi^-$, and $\Xi^- \rightarrow \Lambda \pi^-$. The results will be given using the following definitions of ${\cal A}^{(S)}$ or ${\cal A}^{(P)}$: \begin{eqnarray} {\cal A}^{S,P}_{if}=\alpha^{S,P}_{if} + \left(\beta^{S,P}_{if} - \lambda^{S,P}_{if}\ \alpha^{S,P}_{if}\right){m_K^2 \over 16\pi^2 f_K^2} \log\left({m_K^2 \over \Lambda_\chi^2}\right) \ \ \ , \end{eqnarray} where \begin{eqnarray} \beta^{(P)}_{if}=\beta^{(P1)}_{if}+\beta^{(P2)}_{if} \ \ . \end{eqnarray} The kaon decay parameter and mass are $f_K$ and $m_K$, respectively, and the chiral symmetry breaking scale, $\Lambda_\chi \sim 1$ GeV. The $\alpha^{(S)}_{if}$, $\beta^{(S)}_{if}$ (including both octet and decuplet intermediate states, denoted $\overline{\beta}^{(S)}_{if}$ in Ref.\cite{EJ}), and $\lambda^{(S)}_{if}$ terms can be found in Ref.\cite{EJ}. Despite apparent differences in amplitude definitions, these can be taken straight across because of the units used. The values finally obtained will be different simply because the fit to parameters will include the changes in the P-wave amplitudes. Similarly, the $\alpha^{(P)}_{if}$ and $\lambda^{(P)}_{if}$ are unaffected by the inclusion of the graphs in Figure \ref{P2}. The $\overline{\beta}^{(P)}_{if}$ in Ref.\cite{EJ} will now be called $\beta^{(P1)}_{if}$ and the new graphs will give $\beta^{(P2)}_{if}$. The diagrams in Fig. \ref{P2} yield \begin{eqnarray} \beta^{(P2)}_{\Sigma^+ n} &=& {D \over 3} {h_D +3h_F \over m_\Lambda-m_N} \ \lambda_\Lambda + F \ {h_F -h_D \over m_\Sigma-m_N} \ \lambda_\Sigma - {(F+D)(h_F -h_D) \over m_\Sigma-m_N} \ \lambda_N \nonumber \\ \beta^{(P2)}_{\Sigma^- n} &=& {D \over 3} \ {h_D +3h_F \over m_\Lambda-m_N} \ \lambda_\Lambda - F \ {h_F -h_D \over m_\Sigma-m_N} \ \lambda_\Sigma \nonumber \\ \beta^{(P2)}_{\Lambda p} &=& {2 D \over \sqrt{6}} \ {h_F -h_D \over m_\Sigma-m_N} \ \lambda_\Sigma - {F+D \over \sqrt{6}} \ {3h_F +h_D \over m_\Lambda-m_N} \ \lambda_N \nonumber \\ \beta^{(P2)}_{\Xi \Lambda} &=& - \ {D-F \over \sqrt{6}} \ {3h_F -h_D \over m_\Xi-m_\Lambda} \ \lambda_\Xi + {2 D \over \sqrt{6}} \ {h_F +h_D \over m_\Xi - m_\Sigma} \ \lambda_\Sigma \ \ \ , \end{eqnarray} with \begin{eqnarray} \lambda_N &=& {17 \over 6}D^2-5DF+{15 \over 2}F^2+ {1 \over 2}{\cal C}^2 \nonumber \\ \lambda_\Lambda &=&{7 \over 3}D^2+9F^2+{\cal C}^2 \nonumber \\ \lambda_\Sigma &=& {13 \over 3}D^2+3F^2+{7 \over 3}{\cal C}^2 \nonumber \\ \lambda_\Xi &=& {17 \over 6}D^2+5FD+{15 \over 2}F^2+ {13 \over 6}{\cal C}^2 \ \ . \end{eqnarray} \section{Discussion} The amplitudes obtained from including the diagrams in Fig. \ref{P2} are shown in Tables 1 and 2. The experimental measurements, including errors, are shown in the first column. The second column contains the tree level SU(3) predictions, where the chiral parameters used are the ones extracted from tree-level fits. The third column shows the results of Ref. \cite{EJ}. The fourth column contains the chiral one-loop predictions using weak parameters obtained by fitting only to the S-wave experimental values. To most closely match the analysis of Ref. \cite{EJ}, the strong interaction couplings are chosen to be F=0.4, D=0.61, $|{\cal C}|=1.6$, and ${\cal H}=-1.9$. Letting the weak parameters $h_D$, $h_F$, and $h_C$ vary, a fit to the S-wave decays yields \cite{min} \begin{eqnarray} h_D = -0.32 \pm 0.01, \hskip 1cm h_F = 0.98 \pm 0.03, \hskip 1cm h_C = -1.37 \pm 0.27 \ \ . \end{eqnarray} The errors shown are only those which arise from the experimental variances. The parameter $h_C$ is not well determined and large variations in its value do not appreciably change the predicted amplitudes. The S-wave predictions are essentially unchanged using the parameters above, and the loop corrected chiral predictions are in excellent agreement with experiment, as demonstrated in Ref. \cite{EJ}. The situation for the P-wave predictions is improved for $\Sigma^+ \rightarrow n \pi^+$, where the agreement is within the allowed 30 percent variation for chiral predictions. For the decays $\Sigma \rightarrow n \pi^-$ and $\Lambda \rightarrow p \pi^-$, the additional graphs bring the prediction back to tree level values, while the $\Xi \rightarrow \Lambda \pi^-$ decay remains essentially unchanged. The fifth column in Tables 1 and 2 contains the results from using both S-wave and P-wave amplitudes to fit the weak chiral parameters. The tree level $\Omega$ decays for which there are experimental results are used as well. Expressions for these are in Ref. \cite{EJ}. The strong decays of the decuplet favor midpoint values for $|{\cal C}|$ and ${\cal H}$ of 1.2 and --2.2, respectively \cite{bss}. A fit to $h_D$, $h_F$, and $h_C$ in this scenario yields \begin{eqnarray} h_D = -0.38 \pm 0.01, \hskip 1cm h_F = 0.92 \pm 0.01, \hskip 1cm h_C = 0.74 \pm 0.18 \ \ . \end{eqnarray} The S-waves are still within 30 percent, but the P-waves get worse. The nonleptonic hyperon decays clearly favor a larger value for $|{\cal C}|$ than do the strong decuplet decays. The dependence on ${\cal H}$ is not as sensitive. Using the eight independent nonleptonic hyperon decays, along with the $\Omega$ decays, and ${\cal H}=-2.2$, the parameters ${\cal C}$, $h_D$, $h_F$, and $h_C$ are allowed to vary. The best fit is obtained when \begin{eqnarray} |{\cal C}| = 1.76\pm 0.01&\hskip 2cm& h_D = -0.42\pm 0.01, \nonumber \\ h_F = 0.76\pm 0.01&\hskip 2cm& h_C=0.26\pm 0.10 \end{eqnarray} The matrix of correlation coefficients for this fit, given in the order ($h_C$, $h_D$, $h_F$, ${\cal C}$) is \begin{eqnarray} \left( \begin{array}{rrrr} 1.000&-0.914&-0.974& 0.149 \\ -0.914& 1.000 &0.919&-0.024 \\ -0.974& 0.919& 1.000&-0.009 \\ 0.149&-0.024&-0.009& 1.000 \\ \end{array} \right) \end{eqnarray} The S-wave and P-wave amplitude predictions using these parameters are given in the final column of each Table. The S-waves remain well described, and all but the $\Lambda \rightarrow p \pi^-$ P-wave modes do as well as the S-waves. This later decay amplitude becomes positive for parameter values still within ranges allowed by other observables, but the agreement remains poor. Still, the additional diagrams have improved the situation to the point where the chiral expansion appears to be on more solid footing with respect to the P-wave decays. As Jenkins points out in Ref. \cite{EJ}, the large corrections which the loop diagrams give to the tree-level results need not be taken as evidence that the chiral expansion is ill-behaved if it is the leading order terms which are anomalously small rather than the loop effects which are unnaturally large. \vskip 1.0cm \begin{tabular}{|| c || c | c | c | c | c | c ||} \hline\hline \rule{0cm}{0.5cm} & \multicolumn{6}{c||}{\em S-waves} \\*[0.1cm] \cline{2-7} \rule{0cm}{0.7cm} {\em decay} & exp & tree & theory\cite{EJ} & theory (S) & theory ($\Delta$) & theory \\*[0.1cm] \hline \rule{0cm}{0.5cm}$\hspace{0.2cm} \Sigma^+ \rightarrow n \pi^+$ \hspace{0.2cm} &\hspace{0.2cm} 0.06 $\pm$ 0.01 \hspace{0.2cm} &\hspace{0.2cm} 0.00 \hspace{0.2cm} &\hspace{0.2cm} --0.09 \hspace{0.2cm} &\hspace{0.2cm} --0.09 \hspace{0.2cm} &\hspace{0.2cm} 0.00 \hspace{0.2cm} &\hspace{0.2cm} --0.13 \hspace{0.2cm} \\*[0.1cm] \hline \rule{0cm}{0.5cm}$\hspace{0.2cm} \Sigma^- \rightarrow n \pi^-$ \hspace{0.2cm} &\hspace{0.2cm} 1.88$\pm$0.01 \hspace{0.2cm} &\hspace{0.2cm} 1.21 \hspace{0.2cm} &\hspace{0.2cm} 1.90 \hspace{0.2cm} &\hspace{0.2cm} 1.88 \hspace{0.2cm} &\hspace{0.2cm} 1.74 \hspace{0.2cm} &\hspace{0.2cm} 1.90 \hspace{0.2cm} \\*[0.1cm] \hline \rule{0cm}{0.5cm}$\hspace{0.2cm} \Lambda \rightarrow p \pi^-$ \hspace{0.2cm} &\hspace{0.2cm} 1.42$\pm$0.01 \hspace{0.2cm} &\hspace{0.2cm} 0.91 \hspace{0.2cm} &\hspace{0.2cm} 1.44 \hspace{0.2cm} &\hspace{0.2cm} 1.42 \hspace{0.2cm} &\hspace{0.2cm} 1.44 \hspace{0.2cm} &\hspace{0.2cm} 1.28 \hspace{0.2cm} \\*[0.1cm] \hline \rule{0cm}{0.5cm}$\hspace{0.2cm} \Xi^- \rightarrow \Lambda \pi^-$ \hspace{0.2cm} &\hspace{0.2cm} --1.98$\pm$0.01 \hspace{0.2cm} &\hspace{0.2cm} --1.19 \hspace{0.2cm} &\hspace{0.2cm} --2.04 \hspace{0.2cm} &\hspace{0.2cm} --1.98 \hspace{0.2cm} &\hspace{0.2cm} --1.91 \hspace{0.2cm} &\hspace{0.2cm} --2.02 \hspace{0.2cm} \\*[0.1cm] \hline\hline \end{tabular} \vskip 0.5cm \parbox{6in}{Table 1. The S-wave $\Delta s=1$ hyperon amplitudes. The first column is the experimental result and the next is the tree level prediction of chiral perturbation theory \cite{Wise,EJ}. The third column contains the loop corrected results of Ref.\cite{EJ}. The ``theory (S)'' column gives the fit using S-wave predictions only, with the Ref. \cite{EJ} values $|{\cal C}|$ = 1.6 and ${\cal H}$ = --1.9. The ``theory ($\Delta$)'' column fits both S-wave and P-wave expressions, but uses $|{\cal C}|$=1.2 and ${\cal H}$ = --2.2 taken from the strong decuplet decays. The last column uses the parameters which were obtained from a best fit including both S-waves and P-waves, ${\cal H}$ = --2.2, and ${\cal C}$ fit, including the diagrams of Fig. \ref{P2}.} \vfill\eject \vskip 1.0cm \begin{tabular}{|| c || c | c | c | c | c | c ||} \hline\hline \rule{0cm}{0.5cm} & \multicolumn{6}{c||}{\em P-waves} \\*[0.1cm] \cline{2-7} \rule{0cm}{0.7cm} {\em decay} & exp & tree & theory\cite{EJ} & theory (S) & theory ($\Delta$) & theory \\*[0.1cm] \hline \rule{0cm}{0.5cm}$\hspace{0.5cm} \Sigma^+ \rightarrow n \pi^+$ \hspace{0.2cm} &\hspace{0.2cm} 1.81$\pm$0.01 \hspace{0.2cm} &\hspace{0.2cm} --0.06 \hspace{0.2cm} &\hspace{0.2cm} 0.82 \hspace{0.2cm} &\hspace{0.2cm} 1.54 \hspace{0.2cm} &\hspace{0.2cm} 1.10 \hspace{0.2cm} &\hspace{0.2cm} 1.83 \hspace{0.2cm} \\*[0.1cm] \hline \rule{0cm}{0.5cm}$\hspace{0.2cm} \Sigma^- \rightarrow n \pi^-$ \hspace{0.2cm} &\hspace{0.2cm} --0.06$\pm$0.01 \hspace{0.2cm} &\hspace{0.2cm} 0.13 \hspace{0.2cm} &\hspace{0.2cm} 0.34 \hspace{0.2cm} &\hspace{0.2cm} 0.16 \hspace{0.2cm} &\hspace{0.2cm} 0.34 \hspace{0.2cm} &\hspace{0.2cm} --0.04 \hspace{0.2cm} \\*[0.1cm] \hline \rule{0cm}{0.5cm}$\hspace{0.2cm} \Lambda \rightarrow p \pi^-$ \hspace{0.2cm} &\hspace{0.2cm} 0.52$\pm$0.02 \hspace{0.2cm} &\hspace{0.2cm} --0.28 \hspace{0.2cm} &\hspace{0.2cm} --0.52 \hspace{0.2cm} &\hspace{0.2cm} --0.27 \hspace{0.2cm} &\hspace{0.2cm} --0.51 \hspace{0.2cm} &\hspace{0.2cm} --0.11 \hspace{0.2cm} \\*[0.1cm] \hline \rule{0cm}{0.5cm}$\hspace{0.2cm} \Xi^- \rightarrow \Lambda \pi^-$ \hspace{0.2cm} &\hspace{0.2cm} 0.48$\pm$0.02 \hspace{0.2cm} &\hspace{0.2cm} 0.11 \hspace{0.2cm} &\hspace{0.2cm} 0.35 \hspace{0.2cm} &\hspace{0.2cm} 0.34 \hspace{0.2cm} &\hspace{0.2cm} 0.67 \hspace{0.2cm} &\hspace{0.2cm} 0.48 \hspace{0.2cm} \\*[0.1cm] \hline\hline \end{tabular} \vskip 0.5cm \parbox{6in}{Table 2. The P-wave $\Delta s=1$ hyperon amplitudes. The first column is the experimental result and the next is the tree level prediction of chiral perturbation theory \cite{Wise,EJ}. The third column contains the loop corrected results of Ref.\cite{EJ}. The column labelled ``theory (S)'' uses the parameters obtained from fitting to the S-wave expressions only, with $|{\cal C}|$ = 1.6 and ${\cal H}$ = --1.9, and includes the P-wave diagrams of Fig. \ref{P2}. The ``theory ($\Delta$)'' column fits both S-wave and P-wave decays, but uses $|{\cal C}|$=1.2 and ${\cal H}$ = --2.2 taken from strong decuplet decays. The last column is the result of parameters extracted from a best fit of both S-wave and P-wave expressions, with ${\cal H}$ = --2.2, and ${\cal C}$ fit, including the diagrams of Fig. \ref{P2}.} \vskip 1.0cm \section{Acknowledgements} I would like to thank the Institute for Nuclear Theory at the University of Washington, where much of this work was completed, for their kind hospitality. I gratefully acknowledge advice from Martin Savage and Ted Allen. I thank Martin for many discussions and his always interesting observations and suggestions, and Ted for invaluable assistance with computers, codes, and error analysis. This work is supported in part by the US Dept. of Energy under grant number DE-FG05-90ER40592.
\section{Introduction} The density of states (DOS) in two-dimensional electron systems with a pseudo-gap is a subject of interest for a number of physical situations discussed recently \cite{fradkin,lee,hats,osh,fifra,lud,zie0,tsvel,xiang}. A typical model with a pseudo-gap is represented by the Dirac Hamiltonian in two dimensions \begin{equation} H=i\nabla_1\sigma_1+i\nabla_2\sigma_2+m\sigma_3, \end{equation} where $\sigma_\mu$ are Pauli matrices including the $2\times2$ unit matrix $\sigma_0$. The Dirac equation for a state $\psi$ then is $-\partial\psi/\partial t=H\psi$. The dispersion relation is $E(k_1,k_2)=\pm\sqrt{m^2+k_1^2+k_2^2}$ in the continuum limit. (The lattice will be considered later.) The two signs describe the particle and the hole band, respectively. Both bands touch each other if the Dirac mass vanishes, as one can see in the DOS $\rho(E)\propto |E|\Theta(E^2-m^2)$, where $\Theta$ is the step function. The touching bands is also a feature of a second order phase transition because the decay length of the corresponding Green's function diverges as one goes to the special (critical) point $m=0$. This behavior is indeed formally related to a number of critical phenomena in two-dimensional systems like the ferromagnetic phase transition of the two-dimensional Ising model \cite{dotsenko}. Another physical example, described by the Dirac Hamiltonian, is the degenerate semiconductor which exists for $m=0$ \cite{fradkin}. Furthermore, the large scale limit of a two-dimensional electron gas on a square lattice near the integer quantum Hall transition for certain commensurate flux situations (e.g., half a flux quantum per plaquette) is described by Dirac fermions \cite{hats,osh,fifra,lud,zie0}. A common feature of all these systems is that the DOS at the touching bands (i.e., at $m=0$) is zero, i.e., there is a pseudo-gap. This raises the question whether there is a mechanism which creates states in the pseudo-gap, for instance, electron-electron interaction or quenched disorder. This is important in order to understand if there is a non-vanishing density of low-energy excitations created by interaction or disorder. In this article only the effect of quenched disorder will be analyzed. There is a number of studies for the effect of disorder in the pseudo-gap of Dirac fermions. A numerical calculation for an electron on a square lattice with half a flux quantum per plaquette shows a non-zero density at low energies \cite{hats}. A coherent potential approximation (CPA) of the Dirac fermions with random energy term $E\sigma_0$ added to $H$ also gives a non-zero DOS \cite{fradkin}. A similar result was found for a random mass term $m\sigma_3$ in a modified model with $N$ fermion levels per site, using the $N\to\infty$ limit \cite{zie00}. However, these are essentially mean-field results which may be affected strongly by fluctuations in the two-dimensional system. It is possible that the CPA or $N\to\infty$ result are destroyed by fluctuations in $d=2$. Therefore, as an alternative approach a renormalization group treatment was applied to this problem \cite{dotsenko}. From this it turned out that there is asymptotic freedom indicating that the pseudo-gap, which is controlled by large scale degrees of freedom, is not affected by a random Dirac mass (marginally irrelevant perturbation). However, a rigorous estimation leads to a non-zero lower bound of the DOS in the pseudo-gap \cite{zie1}, at least for a random Dirac mass. The renormalization group calculation indicates that the random energy term is a relevant perturbation, in agreement with the CPA result. A third type of disorder was studied recently by adding a random vector potential to $H$ \cite{lud}. The renormalization group and bosonization treatment indicate a more complicated behavior of the pseudo-gap in this case: The average DOS vanishes like $|E|^\alpha$ with a non-universal exponent $\alpha>0$ if the randomness is weaker than a critical strength. On the other hand, the average DOS diverges if the randomness is stronger than the critical strength because of $\alpha<0$. (For more detailed results see Sect. III.) A similar system with a pseudo-gap is the d-wave superconductor. Nersesyan et al. \cite{tsvel} analyzed this system in $d=2$ and found for the pseudo-gap of the average DOS $\rho(E)\sim |E|^{1/7}$ in contrast to the linear behavior of the pure system. However, this result is in disagreement with others which also find a destruction of the pseudo-gap \cite{xiang}. The effect of disorder in the d-wave superconductor will be discussed in a separate article. The aim of this article is to present an exact solution for the imaginary part of the single particle Green's function of disordered Dirac fermions. \noindent There are several examples in the theory of a quantum particle in a random potential where the average one-particle Green's function can be calculated exactly. Apart from a number of one-dimensional examples \cite{LGP}, there is the Lloyd model \cite{Lloyd}. It is defined by the Hamiltonian $H=H_0+V$, where $H_0$ is a Hermitean matrix (e.g., a tight-binding Hamiltonian for a particles on a lattice). $V$ is a random potential distributed according to a Lorentzian (or Cauchy) distribution \begin{equation} P(V)dV=(\tau/\pi)[\tau^2+(V-V_0)^2]^{-1}dV \label{1} \end{equation} The distribution density has two poles: $V=V_0\pm i\tau$. The Green's function $G(z)=(H+z)^{-1}$ must be averaged with respect to the random potential. $G(z)$ as a function of $V_x$ for a fixed site $x$ is analytic in the upper (lower) complex half-plane if $\Im z$ is positive (negative), respectively. Therefore, the path of integration of $V_x$ can be closed in that half-plane where $G$ is analytic. As a result, only the pole of the Lorentzian density contributes to the integral $\int G P(V_x)dV_x$. This integration can be performed for all lattice sites leading eventually to the average Green's function \begin{equation}\langle G(z)\rangle =[H_0+V_0+z+i sign(\Im z)\tau]^{-1} \equiv G(z+i sign(\Im z)\tau). \label{2} \end{equation} The average DOS then reads \begin{equation}\langle\rho(E)\rangle =-(1/N\pi)\lim_{\epsilon\to0}\Im Tr \langle G(E+i\epsilon)\rangle ,\label{3} \end{equation} where $Tr$ is the trace operator and $N$ is the number of lattice sites. Another examples for an exact solution is the DOS of a particle in a homogeneous magnetic field in two dimensions. If the corresponding Hilbert space of the particle is projected onto the lowest Landau level \cite{Weg}, the average DOS for a white noise potential can be calculated exactly by summing up all terms of the perturbation theory with respect to the white noise potential. The exact solution is related to the fact that the lowest Landau level system is equivalent to a zero-dimensional model. It was discovered by Br\'ezin et al. \cite{BGI} that the latter is a manifestation of the dimensional reduction of the two-dimensional system by 2 due to a supersymmetry of the lowest Landau level problem. Unfortunately, the simplicity of the average DOS of the lowest Landau level cannot be extended to higher Landau levels. It is also in sharp contrast to the complexity of the description of the localization properties \cite{Pruis}. \noindent There is some hope that the treatment of an electron on the square lattice in a strong magnetic field is simpler than a continuum model. The lattice model is motivated by numerical simulations \cite{chalk} and analytic calculations \cite{osh,fifra,lud,zie0}. The reason for a simplification is that the electron near a quantum Hall transition behaves like a Dirac fermion \cite{osh,fifra} because the excitations near the Fermi energy have a linear dispersion. The Hamiltonian of the Dirac fermions on a square lattice with unit lattice constant is \begin{equation} H+i\epsilon\sigma_0=(i\nabla_1+a)\sigma_1+i\nabla_2\sigma_2+m\sigma_3 +i\epsilon\sigma_0. \label{4} \end{equation} The lattice gradient $i\nabla_\mu$, with $\nabla_\mu f(x)=(1/2)[f(x+e_\mu)-f(x-e_\mu)]$ and lattice unit vectors $e_1$, $e_2$, is Hermitean. Two types of disorder are discussed subsequently: a random Dirac mass $m$ and a random vector potential $a$. The vector potential term is chosen in (\ref{4}) in the same way as in Ref.\cite{lud}. It can be considered as a weak disorder approximation of a fluctuating Peierls phase factor in Landau gauge. The Green's function now reads $G(m,i\epsilon)=(H+i\epsilon\sigma_0)^{-1}$, i.e., $m$ and $\epsilon$ correspond with the real and imaginary part of $z$ in the Green's function of the Lloyd model, respectively. The treatment of this problem is rather technical, although the aim is always to find a $G$ with $\Im G=\Im G'$, where the analytic properties of $G$ and $G'$ are different: $G=(H+i\epsilon\sigma_0)^{-1}$, as a function of a random variable at a given site, has poles on {\it both} complex half-planes whereas $G'=(H'+i\epsilon\sigma_0)^{-1}$ has only a pole on {\it one} of the complex half-plane. The Hamiltonian $H'$ is obtained from $H$ by multiplication with a diagonal matrix. The latter depends on the specific type of randomness. The article is organized as follows. In Sect. II the random Dirac mass and in Sect. III a random vector potential are analyzed. The problem of species multiplication due to the lattice is discussed in Sect. IV and the projection onto the homogeneous modes on the lattice is given. \section{Random Dirac Mass} The matrix $H+i\epsilon\sigma_0$ depends on the two complex variables $\pm m_x+i\epsilon$. Thus, in contrast to the Lloyd model, $G(m,i\epsilon)$ may have singularities in both complex half-planes. Therefore, the Green's function of Dirac fermions is similar to the two-particle Green's function of a non-relativistic particle. However, it will be shown subsequently that there is an alternative representation for the imaginary part of the Green's function which depends only on a single complex variable like the Green's function of the Lloyd model. As a first step, $H+i\epsilon\sigma_0$ is multiplied by a diagonal matrix $D\sigma_3$ from the right ($D$ is the staggered diagonal matrix $D_{x,x'}=(-1)^{x_1+x_2} \delta_{x,x'}$ with the two-dimensional space coordinates $x=(x_1,x_2)$) \begin{equation} H'=i(i\nabla_1)D\sigma_2+i(i\nabla_2)D\sigma_1+mD\sigma_0+i\epsilon D\sigma_3, \label{5}\end{equation} where $\nabla_\mu D$ is Hermitean, since $D$ anticommutes with $\nabla_\mu$. Hermitean conjugation of $H'$ yields \begin{equation}{H'}^\dagger=i(i\nabla_1)D\sigma_2+i(i\nabla_2)D\sigma_1+mD \sigma_0-i\epsilon D\sigma_3. \label{6} \end{equation} Moreover, $\sigma_3$ anticommutes with $\sigma_1$ and $\sigma_2$. Consequently, $i\epsilon D\sigma_3$ {\it commutes} with all other terms in $H'$. These properties lead to the product \begin{equation}H'{H'}^\dagger=[i(i\nabla_1)D\sigma_2+i(i\nabla_2)D\sigma_1+mD \sigma_0]^2+\epsilon^2\sigma_0. \label{7} \end{equation} From the definition of $H'$ follows directly \begin{equation}H'{H'}^\dagger= [H+i\epsilon\sigma_0]\sigma_3D\sigma_3D[H+i\epsilon\sigma_0]^\dagger =[H+i\epsilon\sigma_0][H-i\epsilon\sigma_0]= [H-i\epsilon\sigma_0][H+ i\epsilon\sigma_0]. \label{8} \end{equation} The r.h.s. of (\ref{7}) can also be written $H''{H''}^\dagger$ with \begin{equation} H''=i(i\nabla_1)D\sigma_2+i(i\nabla_2)D\sigma_1+(mD+i\epsilon)\sigma_0. \label{9} \end{equation} As a result, $H''$ depends only on one complex variable $(-1)^{x_1+x_2}m_x +i\epsilon$ for a given site $x$. The imaginary part of the Green's function $(H+i\epsilon\sigma_0)^{-1}$ reads $(i/2)([H+i\epsilon\sigma_0]^{-1}-[H-i\epsilon\sigma_0]^{-1})= \epsilon({[H-i\epsilon\sigma_0][H+i\epsilon\sigma_0]})^{-1}$, i.e., it depends on the Hamiltonian only via $H^2$. Therefore, the identity $H^2+i\epsilon^2\sigma_0=H''{H''}^\dagger$ can be used to write \begin{eqnarray} {i\over2}([H+i\epsilon\sigma_0]^{-1}-[H-i\epsilon\sigma_0]^{-1}) =\epsilon({[H-i\epsilon\sigma_0][H+i\epsilon\sigma_0]})^{-1} =\epsilon(H''^\dagger H'')^{-1} \nonumber\\ ={i\over2}[H''^{-1}-(H''^\dagger)^{-1}]. \label{main} \end{eqnarray} Thus the imaginary part of the average Green's function $(H-i\epsilon\sigma_0)^{-1}$ can be calculated exactly for a Lorentzian distribution due to (\ref{main}), where $m_x$ can be integrated out explicitly as in the Lloyd model. Only a pole of the distribution contributes leading to the replacements $\epsilon\to\tau+\epsilon$ and $m\to m_0$ in $H$. This implies \begin{equation} {i\over2}\langle[H+i\epsilon\sigma_0]^{-1}-[H-i\epsilon \sigma_0]^{-1})\rangle ={i\over2}[{\bar H}^{-1}-({\bar H}^\dagger)^{-1}] \label{maina} \end{equation} with \begin{equation} {\bar H}=i\nabla_1\sigma_1+i\nabla_2\sigma_2+i(\epsilon+\tau)\sigma_0+m_0 \sigma_3. \label{13} \end{equation} The imaginary $\tau$-term leads always to an exponential decay of the average Green's function with a typical decay length $\xi\sim (m_0^2+\tau^2)^{-1/2}$ at $\epsilon=0$. Moreover, from equ. (\ref{3}) follows for the average DOS \begin{equation} \langle\rho(i\epsilon,m_0)\rangle=-{1\over N\pi}\Im Tr[{\bar H}^{-1}]. \label{12}\end{equation} The dependence on the energy $E$ is obtained from an analytic continuation $i\epsilon\to i\epsilon+E$. The resulting average DOS is plotted in Fig.1 for $\tau=0.01$ and in Fig.2 for $\tau=0.1$. The non-vanishing DOS is in agreement with a rigorous proof \cite{zie1} and a numerical result \cite{hats}. For Gaussian disorder with variance $g$ there is a lower bound \cite{zie1} \begin{equation} \langle\rho(0,0)\rangle\ge c_1e^{-c_2/g}. \end{equation} with some positive constants $c_1$, $c_2$, independent of $g$. In the continuum limit it was argued, using a one-loop renormalization group calculation, that random fluctuations of the Dirac mass are irrelevant on large scales \cite{lud,tsvel}. This implies a linearly vanishing DOS and a divergent correlation length at $E=m_0=0$. \section{Random Vector Potential} A calculation analogous to that of the random mass system can be performed for a random vector potential. For this purpose the orthogonal transformation $(\sigma_1+\sigma_3)/\sqrt{2}$ is applied to the Hamiltonian \begin{equation} H+i\epsilon\sigma_0\to(i\nabla_1+a)\sigma_3-i\nabla_2\sigma_2+m\sigma_1 +i\epsilon\sigma_0. \label{20} \end{equation} The multiplication of the massless Hamiltonian (i.e., $m=0$) from the r.h.s. with $D'\sigma_3$ (where $D'_{x,x'}=(-1)^{x_2}\delta_{x,x'}$) yields \begin{equation} H'=(i\nabla_1+a)D'\sigma_0-i(i\nabla_2)D'\sigma_1+i\epsilon D'\sigma_3. \label{21} \end{equation} The lattice difference operators $i\nabla_1D'$ and $i(i\nabla_2)D'$ are Hermitean. Since $D'\sigma_3$ commutes with the first two terms of $H'$, one obtains \begin{equation} [H+i\epsilon\sigma_0][H-i\epsilon\sigma_0]=H'H'^\dagger=[(i\nabla_1+a)D' \sigma_0-i(i\nabla_2)D'\sigma_1]^2+ \epsilon^2\sigma_0=H''{H''}^\dagger \label{22} \end{equation} with \begin{equation} H''=(i\nabla_1+a)D'\sigma_0-i(i\nabla_2)D'\sigma_1++i\epsilon\sigma_0 \label{23} \end{equation} which can be used to establish again equation (\ref{main}). For $\langle a\rangle=0$ the average imaginary part of the Green's function and, therefore, the average DOS for $m_0=0$ is related to the Hamiltonian ${\bar H}$ as given in (\ref{13}). In contrast to this result, the bosonization of the Dirac fermions in the continuum limit leads to a different behavior \cite{lud}. For instance, the DOS reads \begin{equation} \langle\rho(E,0)\rangle\sim E^{(2-z)/z}, \end{equation} where $z=1+\Delta_A/\pi$ ($\Delta_A$ is the variance of the fluctuations of the vector potential). I.e., the DOS vanishes at $E=0$ for $z<2$ (weak disorder) and diverges for $z>2$ (strong disorder). The Green's function behaves like \begin{equation} \langle G_{0,x}(E,m=0)\rangle\sim e^{i|x|/\lambda}e^{-|x|/\xi_1} \end{equation} where for $E\sim0$ \begin{equation} \lambda\sim E^{-(1-\Delta_A/\pi)} \end{equation} \begin{equation} \xi_1\sim E^{-1}/\Delta_A. \end{equation} Thus the Green's function decays exponentially for $E\ne0$. There is a critical point $E=0$ where the correlation length diverges with $E^{-1}$. The difference of the results of \cite{lud} and the present work is probably related to the order of taking the continuum limit and averaging over disorder. It is not a consequence of the difference of disorder distributions (Gaussian in \cite{lud} versus Lorentzian distribution here) because the Gaussian distribution could also be treated for ${H''}^{-1}-{H''^\dagger}^{-1}$ in a strong disorder expansion. The result of this expansion is also a finite non-zero DOS and a finite correlation length of $G$. \section{Remark on Species Multiplication} The phenomenon of species multiplication in a fermion lattice theory is well-known from lattice gauge field theories \cite{kogut}. It is due to several nodes in the energy dispersion of the lattice model which indicate the existence of low energy excitations on different length scales. The dispersion of the Dirac fermions considered in this article for $m=a=0$ is $E(k_1,k_2)=\pm\sqrt{\sin^2k_1+\sin^2k_2}$. It has 9 nodes at $k_j=0,\pm\pi$ (cf. Fig.3). In contrast to the lattice model the corresponding continuum model, with $E(k_1,k_2)=\pm\sqrt{k_1^2+k_2^2}$, has low energy excitations only for small wavevectors (i.e., on large scales) as discussed in the Introduction. It will be shown in this Section, using the random mass model of Section II, that the species multiplication is not the reason for the smooth properties of the one-particle Green's function. The degeneracy of the low energy behavior of the lattice model can be lifted by introducing additional terms in the Hamiltonian \cite{kogut}. A possible way is to replace the Hamiltonian $H$ by the new Hamiltonian $H+\delta(\Delta-2)\sigma_3$, where $\delta$ is a positive number ($0<\delta\le1$) and $\Delta$ is a lattice operator with $\Delta f(x)= [f(x+e_1)+f(x-e_1)+ f(x+e_2)+f(x-e_2)]/2$. The dispersion of the new $H$ is $E(k_1,k_2)=\pm\sqrt{\delta^2(\cos k_1+\cos k_2-2)^2+\sin^2k_1+\sin^2k_2}$ for $m=a=0$. This is shown in Fig.4 for $\delta=1/2$. It is not clear to the author which tranformation can be applied to relate the imaginary part Green's function with the new Hamiltonian in order to get the analytic behavior necessary to perform the Cauchy integration with respect to the randomness. However, this difficulty can be circumvented by generalizing $H$ to ${\hat H}$ with \begin{equation} {\hat H}=\pmatrix{ H+\delta(\Delta-2)\sigma_3&m'\sigma_3\cr m'\sigma_3&H-\delta(\Delta+2)\sigma_3\cr }, \end{equation} where $m'$ is a random variable which is statistically independent of $m$ with mean zero. Now the orthogonal transformation \begin{equation} {1\over\sqrt{2}}\pmatrix{ \sigma_0&\sigma_0\cr \sigma_0&-\sigma_0\cr } \end{equation} rotates the diagonal part $(\Delta\sigma_3,-\Delta\sigma_3)$ in the off-diagonal positions and the off-diagonal part into the diagonal position $(m'\sigma_3,-m'\sigma_3)$ such that \begin{equation} {\hat H}=\pmatrix{ H-(2\delta-m')\sigma_3&\Delta\sigma_3\cr \Delta\sigma_3&H-(2\delta+m')\sigma_3\cr }. \end{equation} The random variables $M_x\equiv -2\delta+m_x+m_x'$ and $ M'_x\equiv -2\delta+m_x-m_x'$ in the diagonal part of ${\hat H}$ can now be considered as new independent random variables. The transformation \begin{eqnarray} {\hat H}\to\pmatrix{ \sigma_0&0\cr 0&-\sigma_0\cr }{\hat H}\pmatrix{ D\sigma_3&0\cr 0&D\sigma_3\cr } \nonumber\\ =\pmatrix{ HD\sigma_3-(2\delta-m')D\sigma_0&\Delta D\sigma_0\cr D\Delta\sigma_0&-HD\sigma_3-(2\delta+m')D\sigma_0\cr }={\hat H}' \label{trans} \end{eqnarray} generates the Hermitean matrix ${\hat H}'$. Using for the r.h.s. of (\ref{trans}) the notation $T_0{\hat H}T_1$ and applying the property \begin{equation} T_0{\hat H}T_1=T_1{\hat H}T_0 \label{comm} \end{equation} one obtains \begin{equation} ({\hat H}'-i\epsilon D\gamma_3)({\hat H}'+i\epsilon D\gamma_3)= T_0({\hat H}-i\epsilon\gamma_0)T_1T_0({\hat H}+i\epsilon\gamma_0)T_1 =T_0({\hat H}-i\epsilon\gamma_0)({\hat H}+i\epsilon\gamma_0)T_0 \label{prod1} \end{equation} with the diagonal matrices $\gamma_0=(\sigma_0,\sigma_0)$ and $\gamma_3=(\sigma_3,-\sigma_3)$. Moreover, one has for the imaginary part of the one-particle Green's function as before $(i/2)\Big[({\hat H}+i\epsilon\gamma_0)^{-1} -({\hat H}-i\epsilon\gamma_0)^{-1}\Big] =\epsilon\Big[({\hat H}-i\epsilon\gamma_0)({\hat H} +i\epsilon\gamma_0)\Big]^{-1}$. Due to (\ref{prod1}) and $T_0^{-1}=T_0$ this can be rewritten as \begin{equation} \epsilon T_0\Big[({\hat H}'-i\epsilon D\gamma_3)({\hat H}'+i\epsilon D\gamma_3)\Big] ^{-1}T_0. \label{prod2} \end{equation} The l.h.s. of (\ref{prod1}) reads \begin{equation} ({\hat H}'-i\epsilon D\gamma_3)({\hat H}'+i\epsilon D\gamma_3)= ({\hat H}')^2+\epsilon^2\gamma_0= ({\hat H}'-i\epsilon\gamma_0)({\hat H}'+i\epsilon\gamma_0) \end{equation} because $H'$ and $D\gamma_3$ commute. This implies for (\ref{prod2}) \begin{equation} \epsilon T_0\Big[({\hat H}'-i\epsilon\gamma_0)({\hat H}'+i\epsilon\gamma_0)\Big] ^{-1}T_0 ={i\over2}\Big[({\hat H}'+i\epsilon\gamma_0)^{-1} -({\hat H}'-i\epsilon\gamma_0)^{-1}\Big]. \end{equation} Consequently, the imaginary part of the Green's function satisfies \begin{equation} {i\over2}\Big[({\hat H}+i\epsilon\gamma_0)^{-1} -({\hat H}-i\epsilon\gamma_0)^{-1}\Big] ={i\over2}T_0\Big[({\hat H}'+i\epsilon\gamma_0)^{-1} -({\hat H}'-i\epsilon\gamma_0)^{-1}\Big]T_0, \end{equation} analogously to (\ref{main}). At a given site $x$ the matrix ${\hat H}'+i\epsilon\gamma_0$ depends on the random variables in the combinations $(-1)^{x_1+x_2}M_x+i\epsilon$ and $(-1)^{x_1+x_2}M'_x+i\epsilon$. Assuming a Lorentzian distribution for $M_x$ and $M'_x$, the integration can be performed again as in Sect. II. As a result the imaginary part of the average Green's function is \begin{equation} \Im\pmatrix{ {\bar H}+\delta(\Delta-2)\sigma_3&0\cr 0&{\bar H}-\delta(\Delta+2)\sigma_3\cr }^{-1}, \end{equation} where ${\bar H}$ is the average Hamiltonian (\ref{13}). Thus the lifting of the degeneracy of the nodes in the dispersion relation does not change the analytic behavior of the average one-particle Green's function. \section{Conclusion} An exact expression for the average imaginary part of the one-particle Green's function and the average DOS of two-dimensional lattice Dirac fermions have been derived for a random Dirac mass and for a random vector potential. We have shown that there is a non-zero DOS due to disorder and there is a finite decay length for the average one-particle Green's function. This implies the creation of a non-vanishing density of low-energy excitations due to disorder in a vicinity of $E=M=0$. These lattice results are in agreement with numerical simulation \cite{hats}. However, they are in disagreement with the results of a renormalization group calculation and a bosonization approach for a continuous system of Dirac fermions \cite{lud,tsvel}, where the DOS vanishes or diverges at $E=M=0$. Moreover, the lattice model does not exhibit the critical properties of the Green's function and the DOS found in the renormalization group calculation and in the bosonization approach. It is possible to take the continuum limit of the lattice model after the averaging over disorder, for instance, in the Hamiltonian (\ref{13}). This, however, does not lead to a critical behavior. It seems that the critical behavior of the DOS is a consequence of taking the continuum limit first and performing the averaging over disorder afterwards. This is plausible because the effect of randomness is much stronger in the continuum due to statistically independent fluctuations on arbitrarily short scales. It is shown in Sect. IV that species multiplication, which is a special effect of the lattice model, is not the reason for the smooth behavior of the average DOS. \noindent Acknowledgement: I am grateful to D. Braak for interesting discussions.
\section{Introduction} \label{sec:intro} The introduction of supersymmetry ({\sc SuSy}) can solve the hierarchy problems in the Standard Model (SM) only if {\sc SuSy} is broken at the TeV scale. This implies that the {\sc SuSy} partners of the known particles should be produced at $e^{\scriptscriptstyle +}e^{\scriptscriptstyle -} $\ and $pp$ collider machines planned for the next years. The possibility of observing the new states depends not only on their production cross sections but also on their particular decays and consequent signatures, that might or might not allow their detection in real experiments. Hence, a complete knowledge of the decay structure and relevant branching ratios (BR's) of the lightest {\sc SuSy} states (the first that could be detected) is crucial for discussing the discovery potential of the different machines. In the Minimal Supersymmetric Standard Model (MSSM) \cite{hk}, among the lightest particles in the {\sc SuSy} spectrum, there are 4 neutralinos (the {\sc SuSy} partners of the neutral electroweak (EW) gauge and Higgs bosons) and 2 charginos (the partners of the charged gauge and Higgs bosons). In most scenarios, apart from the Lightest {\sc SuSy} Particle (LSP), which is in general assumed to be the lightest neutralino ($\tilde{\chi}^{\scriptscriptstyle 0}_1 $) (stable and invisible), the particles that could be first observed at future experiments are the next-to-lightest neutralino ($\tilde{\chi}^{\scriptscriptstyle 0}_2 $) and the light chargino ($\tilde{\chi}^{\scriptscriptstyle \pm}_1 $). In particular, the production of $\tilde{\chi}^{\scriptscriptstyle 0}_1 \tilde{\chi}^{\scriptscriptstyle 0}_2 $ pairs at $e^{\scriptscriptstyle +}e^{\scriptscriptstyle -} $ colliders could allow the study of a wide region of the {\sc SuSy} parameter space \cite{amb-mele}. To this respect, it is crucial to know as well as possible the decay characteristics of the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $, that determine the features of the observed signal. Analytical results for the neutralino decay widths have been thoroughly studied in \hbox{Refs.~}{}\cite{bartl}-\cite{hab-wyl}. Nevertheless, at the present time, a complete phenomenological analysis, that investigates the different kinematical and dynamical features of neutralino decays corresponding to different regions of the {\sc SuSy} parameter space, is still missing to our knowledge. In this paper, we present a comprehensive study of the partial decays widths and BR's (including the radiative decay $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma$ and the decay into a light Higgs $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} $) of the next-to-lightest neutralino in the MSSM. The dependence on all the {\sc SuSy} parameters is carefully considered, and non-trivial behaviours are found when varying the different parameters. We assume the usual MSSM framework \cite{hk}, that is: \\ 1) Minimal content of particles and gauge groups, \\ 2) Unification conditions for gauge couplings, gaugino and scalar masses at the GUT (Grand-Unification Theory) scale, \\ 3) $R$-parity conserved. We also assume that the lightest neutralino is the LSP. All masses and couplings are set by choosing the values of a finite set of parameters at the GUT scale: $m_0$ (the common scalar mass), $m_{1/2}$ (the common gaugino mass), $\mu$ (the {\sc SuSy} Higgs-mixing mass) and $\tan \beta $ (the ratio of vacuum expectation values for the two Higgs doublets). A further parameter, $m_{A^{\scriptscriptstyle 0}} $, is needed to describe the Higgs sector, in case one does not use the constraints coming from the requirement that the radiative electroweak-symmetry breaking take place at the correct scale. In Appendices A and B, we describe the equations that allow to get the complete {\sc SuSy} mass spectrum and couplings starting from the above parameters in a standard approximation. We neglect the possibility of mixing between left and right scalar partners of fermions, that can be relevant in the top-stop sector, since this has a marginal role in our study. As for the Higgs sector (that is composed by two minimal doublets), we include the leading logarithmic radiative corrections to masses and couplings, as described in Appendix B. The present work complements \hbox{Ref.~}{}\cite{amb-mele}, where $\tilde{\chi}^{\scriptscriptstyle 0}_1 \tilde{\chi}^{\scriptscriptstyle 0}_2 $ production rates and signatures have been studied at LEP2, by studying extensively the decay features of the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ for a wide choice of {\sc SuSy} parameters. Particular attention is paid to scenarios that are typical of LEP2 physics. The plan of the paper is the following. In section 2, all the next-to-lightest neutralino decay channels in the MSSM are reviewed. Also, we study contour plots of the neutralino-neutralino and neutralino-chargino mass differences, that are crucial for the analysis of the kinematical features of the decays. In section 3, neutralino BR's are presented in the $(\mu, M_2)$ plane. In section 4, some specific scenarios, that are of interest for LEP2 neutralino searches, are analysed. In section 5, the hypothesis of a light Higgs boson is considered. Finally, in section 6, the radiative decay $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma$ is studied. In Appendices A and B, as anticipated above, the neutralino and chargino mass matrices and the scalar-sector mass spectrum are discussed, respectively. \section{Neutralino-decay classification} \label{sec:class} In the MSSM, four fermionic partners of the neutral components of the SM electroweak gauge and Higgs bosons are predicted: the photino $\tilde{\gamma} $, the Z-ino $\tilde{Z} $ (mixtures of the U(1) $\tilde{B}$ and SU(2) $\tilde{W}_3$ gauginos), and the two higgsinos $\tilde{H}^{\scriptscriptstyle 0}_1 $ and $\tilde{H}^{\scriptscriptstyle 0}_2 $ (partners of the two Higgs-doublet neutral components). In general, this interaction eigenstates mix, their mixing being controlled by a mass matrix $Y$ (see Appendix A). By solving a 4-th degree eigenvalue equation, one can find the expressions of $m_{\tilde{\chi}^0_i} $ ($i=1,\ldots,4$) and of the physical composition of the corresponding eigenstates in terms of the set of independent parameters $\mu$, $M_2$ and $\tan \beta $. Here, we are mainly concerned with the two lightest neutralino states ($i =1,2$). The best direct experimental limits on the $\tilde{\chi}^{\scriptscriptstyle 0}_1 $ and $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ masses exclude the ranges $m_{\tilde{\chi}^0_1} < 20 {\rm\,GeV} $ and $m_{\tilde{\chi}^0_2} < 46 {\rm\,GeV} $, under the assumption that $\tan \beta > 2$, at LEP1. These limits disappear if $\tan \beta > 1.6$ \cite{L3}. At LEP2, due to the smaller relative importance of the $Z^{\scriptscriptstyle 0} $-exchange diagram in the $\tilde{\chi}^{\scriptscriptstyle 0}_1 \tilde{\chi}^{\scriptscriptstyle 0}_2 $ production, different physical components of neutralinos (and not only higgsinos) come into play and the common scalar mass $m_0$ becomes a relevant parameter too. In this framework, in order to put new direct limits on the neutralino masses, one must have a complete knowledge also of the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ decay pattern. In \hbox{Ref.~}{}\cite{amb-mele}, the behaviour of the $\tilde{\chi}^{\scriptscriptstyle 0}_{1,2} $ gaugino and higgsino components is studied in detail in the {\sc SuSy} parameter space. This is crucial also to understand the dynamics of the neutralino decays, since different components are coupled to different particles. For instance, in the tree level decays of neutralinos $\tilde{\chi}^{\scriptscriptstyle 0}_i \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_j f\bar{f} $, there are two main contributions coming from the $Z^{\scriptscriptstyle 0} $ and sfermion exchanges. While the gaugino components couple to the scalars, the higgsino components couple only to the $Z^{\scriptscriptstyle 0} $ boson, with different strength (in the $m_f = 0$ limit). Also the neutralino mass spectrum depends on the same three parameters $\mu$, $M_2$ and $\tan \beta $. A detailed discussion on the $\tilde{\chi}^{\scriptscriptstyle 0}_i $ mass spectrum can be found in \hbox{Refs.~}{}\cite{amb-mele} and \cite{bartl89}. In what follows, we list all the possible next-to-lightest neutralino decays in the MSSM. In general, these channels are valid also for heavier neutralinos, although the possibility of cascade decays can make the decay structure of the heavier neutralinos more complicated. \begin{description} \item[a)] Decay into charged leptons: \begin{equation} \tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 e^{\scriptscriptstyle +}e^{\scriptscriptstyle -} \; , \label{n2ton1ee} \end{equation} or $e^{\scriptscriptstyle \pm} \rightarrow \mu^{\scriptscriptstyle \pm} , \; \tau^{\scriptscriptstyle \pm} $; \item[b)] decay into a neutrino pair: \begin{equation} \tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \nu_{\scriptscriptstyle \ell} \bar{\nu}_{\scriptscriptstyle \ell} \; , \label{n2ton1vv} \end{equation} where $\ell = e, \mu, \tau$; \item[c)] decay into a light-quark pair: \begin{equation} \tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 q\bar{q} \; , \label{n2ton1qq} \end{equation} where $q=u,d,s,c,b$; \item[d)] cascade decay through a real chargino: \begin{equation} \begin{array}{r c l} \tilde{\chi}^{\scriptscriptstyle 0}_2 & \rightarrow & f_1 \bar{f}_1^{\prime} \; \tilde{\chi}^{\scriptscriptstyle \pm}_1 \\ & & \phantom{f_1 \bar{f}_1^{\prime} \;} \:\raisebox{1.3ex}{\rlap{$\vert$}}\!\rightarrow f_2 \bar{f}_2^{\prime} \tilde{\chi}^{\scriptscriptstyle 0}_1 \; , \end{array} \label{eq:cascade} \end{equation} where each pair of fermions $f_i f_i^{\prime}$ in the final state is an isospin doublet of either leptons or light quarks; \item[e)] decay into a light scalar ($h^{\scriptscriptstyle 0} $) or pseudoscalar ($A^{\scriptscriptstyle 0} $) Higgs: \begin{mathletters} \label{ntoh} \begin{eqnarray} \tilde{\chi}^{\scriptscriptstyle 0}_2 & \rightarrow & \tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} \label{n2ton1h} \; , \\ \tilde{\chi}^{\scriptscriptstyle 0}_2 & \rightarrow & \tilde{\chi}^{\scriptscriptstyle 0}_1 A^{\scriptscriptstyle 0} \label{n2ton1A} \; , \end{eqnarray} \end{mathletters} where $h^{\scriptscriptstyle 0} $ and $A^{\scriptscriptstyle 0} $ are part of the MSSM Higgs doublet; \item[f)] radiative decay into a photon: \begin{equation} \tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma \; . \label{n2ton1ph} \end{equation} \end{description} The first three channels occur through either a $Z^{\scriptscriptstyle 0} $ or a scalar-particle exchange. Different scalar partners come into play: (left or right) selectron (in channel {\bf a}), (left) sneutrino (in channel {\bf b}) and (left or right) squark (in channel {\bf c}) (see \hbox{Fig.~}{}\ref{fey3bneut}). Assuming massless fermions makes the channels proceeding through neutral Higgs bosons vanish in {\bf a}, {\bf b} and {\bf c}. We name $s$-channels the contributions from diagrams with the two neutralinos entering the same vertex ($Z^{\scriptscriptstyle 0} $ exchange), and ($t,u$)-channels the ones where the two neutralinos enter different vertices (sfermion exchange). Whenever the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ is heavier than some scalar fermions, the corresponding channels will proceed through two steps via real sparticles. A possible gluino in the final state ($\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{g} q\bar{q} $) is excluded by the gaugino mass unification hypothesis, that makes gluinos considerably heavier than light neutralinos. Cascade decays through a real chargino ({\bf d}) occur via similar graphs (\hbox{Fig.~}{}\ref{fey3bcasc}). The diagrams for the second-step decay (\hbox{cfr.}{}\ \hbox{Eq.~}{}(\ref{eq:cascade})) can be obtained by the same graphs by exchanging the neutralino and the chargino. As for the channel {\bf e}, there are five possible Higgses (either neutral or charged) that could contribute to the tree-level $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ decays into a scalar Higgs boson plus a light neutralino or chargino. For the next-to-lightest neutralino, only decays into the two lightest bosons (\hbox{i.e.}{}, the lightest neutral scalar and the pseudoscalar Higgses) can be present, for the moderate $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ masses we are considering here (\hbox{Fig.~}{}\ref{fey2bhiggs}). One important point to keep in mind in the decay study is that, whenever the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ can decay into a real scalar plus a fermion (\hbox{e.g.}{}, a selectron plus an electron or a Higgs plus a lightest neutralino), this channel tends to saturate the corresponding width and BR. The same occurs when the mass difference between the two lightest neutralino is sufficient to allow the decay into a real $Z^{\scriptscriptstyle 0} $. In the latter case, the relevant BR's for different signatures recover the $Z^{\scriptscriptstyle 0} $ ones. However, the last possibility never occurs in the LEP2 parameter regions. We point out that, apart from the decays into Higgses, that are considered only if the two-body on-shell decay is allowed by the phase space, our treatment of the three-body decays always takes into account properly the possibility of decays into two real particles, whenever this is permitted. In \hbox{Fig.~}{}\ref{n2-n1mass}, we show the contour plot for the mass difference between $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ and $\tilde{\chi}^{\scriptscriptstyle 0}_1 $, for $\tan \beta = 1.5$ and 30, in the $(\mu, M_2)$ plane. From these plots, one can immediately infer, for a given scalar mass, which are the parameter regions where the decays into real scalars are kinematically allowed, and consequently can dominate the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ decay. Furthermore, in \hbox{Fig.~}{}\ref{n2-c1mass}, the difference between $m_{\tilde{\chi}^0_2} $ and $m_{\tilde{\chi}^{\pm}_1} $ is plotted. Shaded area represent situations where this difference is negative and the neutralino cascade decays through a chargino are not allowed. For small $\tan \beta $, one can anticipate a sizeable BR for cascade decays in the positive $\mu$ half-plane (\hbox{cfr.}{}\ section 3). Diagrams contributing to the radiative decay $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma$ are shown in \hbox{Fig.~}{}\ref{feyrad}, where the corresponding graphs with clockwise circulating particles in the loops must be added. The fields $G^{\scriptscriptstyle \pm} $ are the Goldstone quanta giving masses to charged vector bosons. We assume the nonlinear R-gauge, that is described in \hbox{Ref.~}{}\cite{hab-wyl}. One can see that there are many physical charged particles flowing in the loops: all charged standard fermions and their corresponding scalar partners, the charged vector and Higgs bosons and their fermionic partners, the charginos. In the {\sc SuSy} parameter scheme we adopt, relevant contributions come mostly from the $W^{\scriptscriptstyle \pm} $/chargino and the top/stop loops, with non-negligible interferences. In this decay, the visible part of the final state is given by a monochromatic photon. {}From \hbox{Fig.~}{}\ref{n2-n1mass}, one can get information on the final photon energy. In the following analysis, we mainly concentrate on the MSSM parameter regions not excluded at LEP1. Particular attention is given to regions explorable at the forthcoming experiments, especially at LEP2 (that are shown in \hbox{Fig.~}{}\ref{lep2scen}). For definiteness, we restrict to the following ranges of {\sc SuSy} parameters: \begin{mathletters} \label{eq:parng} \begin{eqnarray} 0 \le & M_2 & \le 4 M_{\scriptscriptstyle Z} \\ -4M_{\scriptscriptstyle Z} \le & \mu & \le 4 M_{\scriptscriptstyle Z} \\ 40 {\rm\,GeV} \le & m_0 & \le 500 {\rm\,GeV} \\ 1 < & \tan \beta & \le 60 \\ M_{\scriptscriptstyle Z} \le & m_{A^{\scriptscriptstyle 0}} & \le 3 M_{\scriptscriptstyle Z} \end{eqnarray} \end{mathletters} The lower limit on $m_0$ is connected to present experimental limits on the masses of the {\sc SuSy} partners of leptons and quarks. It generally excludes scenarios where the LSP is a scalar. \section{Study of the $\protect\tilde{\chi}^{\scriptscriptstyle 0}_2$ BR's in the $(\mu, M_2)$ plane} \label{sec:BRvsmum2} In this section, we make a detailed study of the BR's for the decay channels {\bf a-e} (defined in the previous section) in the $(\mu, M_2)$ plane, at fixed values of $m_0$ and $\tan \beta $. We assume $m_{A^{\scriptscriptstyle 0}} = 3 M_{\scriptscriptstyle Z} $, that, unless $\tan \beta $ is near 1, generally implies a $h^{\scriptscriptstyle 0} $ mass above the threshold for the channel $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} $, in the $(\mu, M_2)$ region covered by LEP2 searches. The light-Higgs case will be considered in section 5, while in section 6 we will concentrate on the radiative decay $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma$. Everywhere, the widths and BR's connected to the decays into charged leptons are relative to a single species, while the decays into neutrinos are summed over three families and the decays into quarks over five light flavours. Analogously, for cascade decays, the BR for the leptonic channel is for one single species, while the hadronic channels are summed over two light quark doublets. At this stage, the second-step decay of the $\tilde{\chi}^{\scriptscriptstyle \pm}_1 $ is not considered. The BR's for each channel are studied in the $(\mu, M_2)$ plane for four different choices of the $m_0$ and $\tan \beta $ parameter, namely: \begin{mathletters} \label{br_cases} \begin{eqnarray} {(\rm a)}: \:\: (m_0, \tan \beta ) & = & (M_{\scriptscriptstyle Z} , 1.5), \\ {(\rm b)}: \:\: \phantom{(m_0, \tan \beta )} & = & (M_{\scriptscriptstyle Z} , 30), \\ {(\rm c)}: \:\: \phantom{(m_0, \tan \beta )} & = & (3M_{\scriptscriptstyle Z} , 1.5), \\ {(\rm d)}: \:\: \phantom{(m_0, \tan \beta )} & = & (3M_{\scriptscriptstyle Z} , 30), \end{eqnarray} \end{mathletters} for $|\mu| \le 4 M_{\scriptscriptstyle Z} $ and $0 \le M_2 \le 4 M_{\scriptscriptstyle Z} $. The observed behaviour is in general highly non-trivial, due to both the sharp dependence of the neutralino physical composition on $\mu$ and $M_2$ \cite{amb-mele} and the corresponding variation in the neutralino mass spectrum (\hbox{cfr.}{}\ \hbox{Fig.~}{}\ref{n2-n1mass}). Also a weak dependence on $M_2$ comes from the spectrum of the scalar masses that enter the $t$-channel contributions to the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ decay (\hbox{Fig.~}{}\ref{fey3bneut}). In \hbox{Fig.~}{}\ref{brnee}, we present the BR for the decay $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 e^{\scriptscriptstyle +}e^{\scriptscriptstyle -} $. In order to get large BR's, in this case, one needs relatively small $m_0$ values, so that the $t$-channel contributions get substantial with respect to the $Z^{\scriptscriptstyle 0} $-exchange channel. In this way, one obtains leptonic BR's much larger than the corresponding BR($Z^{\scriptscriptstyle 0} \rightarrow \ell^{\scriptscriptstyle +}\ell^{\scriptscriptstyle -} $). For instance, when $m_0 = M_{\scriptscriptstyle Z} $ (\hbox{Fig.~}{}\ref{brnee}a,b), one always gets wide regions of the plane where BR($\tilde{\chi}^{\scriptscriptstyle 0}_1 e^{\scriptscriptstyle +}e^{\scriptscriptstyle -} $) is larger than 25\% (that corresponds to a BR $\ge 75\%$ when summed over three lepton species). However, when considering the LEP2 realm (\hbox{Fig.~}{}\ref{lep2scen}), the relative importance of these regions is reduced, especially for large $\tan \beta $ values. Note that at large $\tan \beta $, the behaviour tends to be more symmetric with respect to the line $\mu = 0$ (\hbox{Fig.~}{}\ref{brnee}b,d). In order to guarantee the detectability of the $\tilde{\chi}^{\scriptscriptstyle 0}_1 e^{\scriptscriptstyle +}e^{\scriptscriptstyle -} $ final state in ordinary collider experiments, it is also useful to consider a threshold on the minimum energy for an observable $e^{\scriptscriptstyle +}e^{\scriptscriptstyle -} $ state. The effect of this condition can be guessed through \hbox{Fig.~}{}\ref{n2-n1mass}, since $(m_{\tilde{\chi}^0_2} - m_{\tilde{\chi}^0_1} )$ is directly connected to the final $e^{\scriptscriptstyle +}e^{\scriptscriptstyle -} $ energy. For instance, one can see that the rejection of the areas where $(m_{\tilde{\chi}^0_2} - m_{\tilde{\chi}^0_1} ) < 20 {\rm\,GeV} $ has a moderate influence on LEP2 physics (\hbox{cfr.}{}\ \hbox{Fig.~}{}\ref{lep2scen}). In \hbox{Fig.~}{}\ref{brnee}a, the large BR at moderate values of $M_2$ and $\mu < 0$ is due to the opening of the tree-level channel $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow e^{\scriptscriptstyle \mp} \tilde{e}^{\scriptscriptstyle \pm}_{\scriptscriptstyle R} $ at $m_0 \approx M_{\scriptscriptstyle Z} $, which is not contrasted by $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \nu_{\scriptscriptstyle e} \tilde{\nu}_{\scriptscriptstyle e,L} $ (in general $m_{\tilde{e}_R} < m_{\tilde{\nu}_{e,L}} $, \hbox{cfr.}{}\ Appendix B). See \hbox{Fig.~}{}\ref{wbrvsm0a} for further details. In \hbox{Fig.~}{}\ref{brnvv}, the channel $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \nu_{\scriptscriptstyle \ell} \bar{\nu}_{\scriptscriptstyle \ell} $ is studied. As in the previous case, low $m_0$ values tend to enhance the BR. Note that, in both the charged lepton and the neutrino case, the corresponding $Z^{\scriptscriptstyle 0} $ BR's are recovered in the region of small $|\mu|$ and $M_2 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\;$ (1-2) $M_{\scriptscriptstyle Z} $. Indeed, in this region, the higgsino components and, consequently, the $Z^{\scriptscriptstyle 0} $-exchange channel are dominant, independently of $m_0$ and $\tan \beta $. This is also true for the hadronic channel that is considered in \hbox{Fig.~}{}\ref{brnqq}. On the other hand, in the hadronic decay, a low $m_0$ value can decrease the BR with respect to the $Z^{\scriptscriptstyle 0} $-channel expectation. This is due, for a given $m_0$, to the larger value of the squark masses (entering the $t$-channel contribution), compared to the slepton masses (\hbox{cfr.}{}\ Appendix B). For large $m_0$, $t$-channels tend to vanish, and the BR for $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 f\bar{f} $ recovers the $Z^{\scriptscriptstyle 0} \rightarrow f\bar{f} $ one. Cascade-decay BR's are shown in \hbox{Figs.~}{}\ref{brclv} and \ref{brcqq} for the channels $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle \pm}_1 e^{\scriptscriptstyle \mp} \nu_{\scriptscriptstyle e} $ and $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \sum_q \tilde{\chi}^{\scriptscriptstyle \pm}_1 q \bar{q}^{\prime}$, respectively. At small values of $\tan \beta $, the importance of this channel is restricted to the positive-$\mu$ half-plane in connection to the regions where $\tilde{\chi}^{\scriptscriptstyle \pm}_1 $ is sufficiently lighter than $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ (\hbox{cfr.}{}\ \hbox{Figs.~}{}\ref{brclv}a,c, \ref{brcqq}a,c and \ref{n2-c1mass}). The leptonic channel can reach a BR of 10\% for each leptonic species at low $\tan \beta $. The relevance of this decay is further increased at larger $m_0$ (\hbox{cfr.}{}\ \hbox{Fig.~}{}\ref{brclv}c). Raising $\tan \beta $ makes the pattern symmetrical with respect to the $\mu = 0$ axis, although the BR never reaches a sizeable level in the interesting regions (\hbox{cfr.}{}\ \hbox{Figs.~}{}\ref{brclv}b,d). An analogous situation is observed for the decay into $\tilde{\chi}^{\scriptscriptstyle \pm}_1 q \bar{q}^{\prime}$ in \hbox{Fig.~}{}\ref{brcqq}. In \hbox{Fig.~}{}\ref{brnlh} we study the channel $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} $. Whenever the light-Higgs mass is lighter than the difference $(m_{\tilde{\chi}^0_2} - m_{\tilde{\chi}^0_1} )$, this process has a large BR due to the two-body nature of the decay. The Higgs mass spectrum is fixed here by $m_{A^{\scriptscriptstyle 0}} = 3 M_{\scriptscriptstyle Z} $, that corresponds to $m_{h^{\scriptscriptstyle 0}} $ in the range $ 50 \div 90 {\rm\,GeV} $, at $\tan \beta = 1.5$ and $100 \div 130 {\rm\,GeV} $, at $\tan \beta = 30$, for the assumed range of $m_0$ and $M_2$ (\hbox{cfr.}{}\ Appendix B). The most favourable case for the process $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} $ is the one with small $\tan \beta $ and $m_0$ (\hbox{cfr.}{}\ \hbox{Fig.~}{}\ref{brnlh}a). Even in this case, the decay threshold opens mostly at relatively large values of $|\mu|$, that correspond to heavier neutralinos. Increasing $m_0$ (\hbox{cfr.}{}\ \hbox{Fig.~}{}\ref{brnlh}c), slightly restricts the allowed regions, due to the $m_{\tilde{t}} $ dependence of the $m_{h^{\scriptscriptstyle 0}} $ radiative corrections (\hbox{cfr.}{}\ Appendix B), but, at the same time, increases the branching fraction, due to the depletion of all the other $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 f\bar{f} $ channels. At $\tan \beta = 30$, the decay $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} $ is not allowed in almost the whole $(\mu, M_2)$ plane considered, due to the increase of $m_{h^{\scriptscriptstyle 0}} $ with $\tan \beta $ (\hbox{cfr.}{}\ Appendix B) and also to the decrease of $(m_{\tilde{\chi}^0_2} - m_{\tilde{\chi}^0_1} )$ (\hbox{cfr.}{}\ \hbox{Fig.~}{}\ref{n2-n1mass}). In \hbox{Figs.~}{}\ref{brnlh}b,d an intermediate-$\tan \beta $ situation is shown. Note that the case of a lighter $h^{\scriptscriptstyle 0} $ (or $m_{A^{\scriptscriptstyle 0}} $) can considerably alter the pattern of the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ BR's. The case of a lighter Higgs will be considered in section 5. \section{Neutralino decays at LEP2} \label{sec:LEP2dec} In our analysis of the production of $\tilde{\chi}^{\scriptscriptstyle 0}_1 \tilde{\chi}^{\scriptscriptstyle 0}_2 $ pairs at LEP2 in \hbox{Ref.~}{}\cite{amb-mele}, we have identified particular regions and scenarios in the parameter space. These are characterized by specific dynamical and kinematical properties of the light neutralinos. We have defined as NR$^{\rm \pm}$ ({\it Neutralino Regions}) the areas of the $(\mu, M_2)$ plane where: \begin{equation} m_{\tilde{\chi}^0_1} + m_{\tilde{\chi}^0_2} < \sqrt{s} < m_{\tilde{\chi}^{\pm}_1} \; , \label{NR} \end{equation} at fixed $\tan \beta $, where $\sqrt{s} $ is the \hbox{c.m.}{}\ collision energy at LEP2 ($\sqrt{s} \simeq 190 {\rm\,GeV} $) (\hbox{Fig.~}{}\ref{lep2scen}). In this regions, chargino pair production is kinematically forbidden, while, for moderate values of scalar masses, the $\tilde{\chi}^{\scriptscriptstyle 0}_1 \tilde{\chi}^{\scriptscriptstyle 0}_2 $ production can have sizeable cross sections. Of course, neutralino-pair production can be of help also below the Neutralino Regions, where it complements chargino production. On the other hand, the largest rates for $e^{\scriptscriptstyle +}e^{\scriptscriptstyle -} \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \tilde{\chi}^{\scriptscriptstyle 0}_2 $ arise for $|\mu| \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$<$}\; M_{\scriptscriptstyle Z} $ and $ M_2 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; M_{\scriptscriptstyle Z} $ (what we call HCS$^{\rm \pm}$, that stands for {\it High Cross Section} regions), although, in these zones, chargino production is also allowed. Here, the higgsino components of $\tilde{\chi}^{\scriptscriptstyle 0}_1 $ and $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ are dominant and the main production mechanism is through $Z^{\scriptscriptstyle 0} $ exchange. In these regions, for $\tan \beta = 1.5$, we have chosen a set of six specific points (shown in \hbox{Fig.~}{}\ref{lep2scen}a), that can schematize the spectrum of possibilities for the neutralino couplings and masses: scenarios A, B, C and D in the Neutralino Regions and H$^{\rm \pm}$ in the High Cross Section regions. In \hbox{Tab.~}{}\ref{tab:scentgb15}, we show the masses and physical components of the two lightest neutralinos and the mass spectrum of charginos and heavier neutralinos, corresponding to these points in the $(\mu, M_2)$ plane. Moreover, we show the sfermion spectrum (that also influences the decay properties of neutralinos) corresponding to these cases for $m_0 = M_{\scriptscriptstyle Z} $. A more detailed analysis of the dynamical characteristics of these scenarios can be found in \hbox{Ref.~}{}\cite{amb-mele}. In this section, we present a study of the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ widths and BR's in these specific cases. Although the values $\tan \beta = 1.5$ (associated to the definition of such scenarios) and $m_0 = M_{\scriptscriptstyle Z} $ correspond to particularly favourable cases for neutralino production rates, we also study the behaviour of decay widths and BR's in a large range of $\tan \beta $ and $m_0$ values. We will call $\tilde{A}$-$\tilde{H}^{\pm}$ the scenarios with the same values of $\mu$ and $M_2$ as A-H$^{\rm \pm}$, but $\tan \beta \ne 1.5$ (see, \hbox{e.g.}{}, \hbox{Fig.~}{}\ref{lep2scen}b where $\tan \beta = 30$). If not otherwise specified, we will assume that decays into real Higgs bosons are not kinematically allowed. In this case, there is some influence of the Higgs sector only in the radiative channel. Accordingly, the results presented in this section are obtained for $m_{A^{\scriptscriptstyle 0}} = 3 M_{\scriptscriptstyle Z} $. The effect of varying $m_{A^{\scriptscriptstyle 0}} $ will be discussed in section 5. In \hbox{Figs.~}{}\ref{wbrvsm0a}--\ref{wbrvsm0hp}, we present all partial widths and BR's versus $m_0$ in the scenarios A, B, C, D, H$^{\rm -}$ and H$^{\rm +}$. In order to understand the general pattern of the decay widths, one must recall the specific ordering of the scalar masses at fixed $m_0$, that is predicted by the renormalization group equations and unification assumptions. For instance, one always gets: $m_{\tilde{q}_{L,R}} > m_{\tilde{\ell}_{L,R}} $ and $m_{\tilde{f}_L} > m_{\tilde{f}_R} $. When $m_0$ is sufficiently small, as to allow $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ decays into one or more real scalars, the largest decay widths are associated to the corresponding channels, that in general are the leptonic ones. For instance, this occurs for $m_0 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$<$}\; 100 {\rm\,GeV} $, in \hbox{Fig.~}{}\ref{wbrvsm0a}, where we study the scenario A ($\tilde{\chi}^{\scriptscriptstyle 0}_{1,2} $ dominantly gauginos, \hbox{cfr.}{}\ \hbox{Tab.~}{}\ref{tab:scentgb15}). One can see that the largest rates (up to $1 \div 100 {\rm\,MeV} $) are by far those corresponding to decays into real sleptons. The width for $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \nu_{\scriptscriptstyle \ell} \bar{\tilde{\nu}}_{\scriptscriptstyle \ell,L} \rightarrow \nu_{\scriptscriptstyle \ell} \bar{\nu}_{\scriptscriptstyle \ell} \tilde{\chi}^{\scriptscriptstyle 0}_1$ is summed over all neutrino flavours, contrary to the charged-lepton channel. For $m_0 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$<$}\; 60 {\rm\,GeV} $, all sleptons are produced on the mass-shell. By increasing $m_0$, one meets, in order, the $m_{\tilde{\ell}_L} $, $m_{\tilde{\nu}_{\ell,L}} $ and $m_{\tilde{\ell}_R} $ thresholds. At large $m_0$, only the $Z^{\scriptscriptstyle 0} $-exchange contributions survive. Note also the peaks of the leptonic cascade-decay width, that correspond to the chain of on-shell decays $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{e}^{\scriptscriptstyle \pm}_{\scriptscriptstyle L} e^{\scriptscriptstyle \mp} \rightarrow \tilde{\chi}^{\scriptscriptstyle \pm}_1 \nu_{\scriptscriptstyle e} e^{\scriptscriptstyle \mp}$ and $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\nu}_{\scriptscriptstyle e,L} \nu_{\scriptscriptstyle e} \rightarrow \tilde{\chi}^{\scriptscriptstyle \pm}_1 e^{\scriptscriptstyle \mp} \nu_{\scriptscriptstyle e}$. When $m_0 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; 100 {\rm\,GeV} $, no two-body decay is allowed. On the other hand, in the parameter range considered, one has in general $m_{\tilde{q}_{L,R}} > m_{\tilde{\chi}^0_2} $ and the width for the $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow q\bar{q} \tilde{\chi}^{\scriptscriptstyle 0}_1 $ is relatively small and almost constant (between 0.1 and 1 KeV) when varying $m_0$. Concerning the radiative decay, the curves are obtained through the complete results of \hbox{Ref.~}{}\cite{hab-wyl}. The width for $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma$ never exceeds 1 KeV for $m_0 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; M_{\scriptscriptstyle Z} $. The corresponding BR pattern closely reflects the scalar-mass threshold structure (\hbox{Fig.~}{}\ref{wbrvsm0a}). Indeed, for $m_0 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$<$}\; 75 {\rm\,GeV} $, the BR for the invisible channel $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \sum_{\ell}\nu_{\scriptscriptstyle \ell} \bar{\nu}_{\scriptscriptstyle \ell} \tilde{\chi}^{\scriptscriptstyle 0}_1 $ is more than 80\%. Hence, in this regime, the next-to-lightest neutralino is phenomenologically almost equivalent to the LSP, which means that most of the times it just produces missing energy and momentum in the final states. This effect tends to concern even larger ranges of $m_0$ when $\tan \beta $ increases, due to the relative lowering of the sneutrino mass with respect to the charged-slepton masses. Also, notice that the presence of two close thresholds for the decay into a real $\tilde{\nu}_{\scriptscriptstyle \ell,L} $ and a real $\tilde{\ell}_{\scriptscriptstyle R} $ gives rise to a peak structure in the BR($\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \ell^{\scriptscriptstyle +}\ell^{\scriptscriptstyle -} \tilde{\chi}^{\scriptscriptstyle 0}_1 $) at $m_0 \simeq 83 {\rm\,GeV} $. This corresponds to a fast deepening in the invisible BR. For $80 {\rm\,GeV} \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$<$}\; m_0 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$<$}\; 200 {\rm\,GeV} $, the largest BR is that for charged leptons ($\;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; 60\%$, for all the three lepton species). For $m_0 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; 200 {\rm\,GeV} $, the hadronic channel gets more and more important. It reaches 80\% for $m_0 \simeq \frac{1}{2} {\rm\,TeV} $. The BR for the radiative decay grows with $m_0$, although at $m_0 \simeq 500 {\rm\,GeV} $ one still has only BR$\simeq 4\%$. Anyhow, concerning searches at LEP2, one has to keep in mind that, for $m_0 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; 300 {\rm\,GeV} $, the production rate for $\tilde{\chi}^{\scriptscriptstyle 0}_1 \tilde{\chi}^{\scriptscriptstyle 0}_2 $ pairs is below the detectability threshold for a realistic machine luminosity (see \hbox{Ref.~}{}\cite{amb-mele}). In \hbox{Fig.~}{}\ref{wbrvsm0b}, we deal with the scenario B, where the $\tilde{\chi}^{\scriptscriptstyle 0}_1 $ is mainly a photino and the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ is mainly a $\tilde{H}^{\scriptscriptstyle 0}_b $ (\hbox{cfr.}{}\ \hbox{Tab.~}{}\ref{tab:scentgb15}). In general, the behaviour of widths and BR's is qualitatively similar to the ones in scenario A, apart from the absence of the decay into $\tilde{\ell}_{\scriptscriptstyle L} $ (whose mass is above threshold) and the presence of rather strong destructive interference effects in the leptonic channels. The latter are clearly visible for the process $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \nu_{\scriptscriptstyle \ell} \bar{\nu}_{\scriptscriptstyle \ell} $ around $m_0 \simeq 64 {\rm\,GeV} $ and in the case $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 e^{\scriptscriptstyle +}e^{\scriptscriptstyle -} $ for $m_0 \simeq 130 {\rm\,GeV} $ (\hbox{Fig.~}{}\ref{wbrvsm0b}). Indeed, the different physical nature of $\tilde{\chi}^{\scriptscriptstyle 0}_1 $ and $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ gives rise, in particular $m_0$ ranges, to a comparable size for the $s$- and $t$-channel decay contributions with negative interference of the same order of magnitude. For instance, in the minimum of the width for the decay into electrons, the $s$-channel and $t$-channel contribute 65 and 50 MeV, respectively, while their interference gives -95 MeV. In the scenario C (\hbox{cfr.}{}\ \hbox{Tab.~}{}\ref{tab:scentgb15}) the larger value of $M_2$ with respect to the previous scenarios generates larger masses in the scalar sector, particularly in the left-handed sector. As a consequence, only the $\tilde{\ell}_{\scriptscriptstyle R} $ can go on mass-shell, while both the invisible and the hadronic widths, that are dominated by the $Z^{\scriptscriptstyle 0} $-exchange, keep constant (\hbox{Fig.~}{}\ref{wbrvsm0c}). As for BR's, the charged leptonic channels saturate the width up to $m_0 \simeq 64 {\rm\,GeV} $ (BR$\simeq 33\%$ for each lepton species) (\hbox{Fig.~}{}\ref{wbrvsm0c}). Note that the relatively small widths for the tree level channels enhance the BR for the radiative decay, for $m_0 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; 70 {\rm\,GeV} $, up to about 15\%. The scenario D is chosen in the positive-$\mu$ range (contrary to the previous ones) and, in particular, in the area of the Neutralino Regions where $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ cascade decays through a light chargino are allowed. The physical composition of $\tilde{\chi}^{\scriptscriptstyle 0}_{1,2} $ is given by a mixture of comparable components of $\tilde{\gamma} $ and $\tilde{Z} $ with a small percentage of higgsino components (\hbox{cfr.}{}\ \hbox{Tab.~}{}\ref{tab:scentgb15}). In \hbox{Fig.~}{}\ref{wbrvsm0d}, the only qualitative new feature in the decay pattern is the presence of sizeable widths for cascade decays. The latter are almost constant versus $m_0$ and give rise to BR's up to 15\% for the channel $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle \pm}_1 q \bar{q}^{\prime}$ and up to 3\% for $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle \pm}_1 e^{\scriptscriptstyle \pm} \nu_{\scriptscriptstyle \ell} $ at $m_0 \simeq 500 {\rm\,GeV} $. In \hbox{Figs.~}{}\ref{wbrvsm0hm} and \ref{wbrvsm0hp}, we present the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ widths and BR's in the High Cross Section regions. In particular, we consider the scenarios H$^{\rm \pm}$, defined in \hbox{Tab.~}{}\ref{tab:scentgb15}. One can check that the higgsino components in these cases are dominant and the $\tilde{\chi}^{\scriptscriptstyle 0}_1 $ and $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ are mostly coupled to the $Z^{\scriptscriptstyle 0} $ boson. Hence, a small dependence on the scalar masses is found. This is the case especially in the H$^{\rm -}$ scenario (\hbox{cfr.}{}\ \hbox{Fig.~}{}\ref{wbrvsm0hm}), where the BR for the $q\bar{q} $, $\ell^{\scriptscriptstyle +}\ell^{\scriptscriptstyle -} $ and $\nu_{\scriptscriptstyle \ell} \bar{\nu}_{\scriptscriptstyle \ell} $ channels are quite the same as for $Z^{\scriptscriptstyle 0} $ decays. In \hbox{Figs.~}{}\ref{wbrvsm0hm} and \ref{wbrvsm0hp}, the widths and BR's for the cascade channels are also reported. The relative importance of these decay modes is considerable only in the H$^{\rm +}$ case, where the corresponding BR's reach about 18\% for the hadronic mode, and more than 2\% for each leptonic channel. We now proceed to the study of the $\tan \beta $ dependence of the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ decay pattern. To this end, as anticipated, we define six new scenarios $\tilde{A}$, $\tilde{B}$, $\tilde{C}$, $\tilde{D}$, $\tilde{H}^{\pm}$, that are obtained from the above scenarios by fixing $m_0 = M_{\scriptscriptstyle Z} $, and letting free the $\tan \beta $ value. We then study the effect of changing $\tan \beta $ in the range $1 \div 60$. Note that, although scenarios A-D were originally defined as lying in the Neutralino Regions, the variation of $\tan \beta $ can shift such regions above some of these scenarios (\hbox{cfr.}{}\ \hbox{Fig.~}{}\ref{lep2scen}). This corresponds to study situations in the $(\mu, M_2)$ plane where also chargino production is allowed. Furthermore, changing $\tan \beta $ varies both the mass spectrum and the physical composition of $\tilde{\chi}^{\scriptscriptstyle 0}_{1,2} $. For instance, in \hbox{Fig.~}{}\ref{massdiffvstgb}, the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $-$\tilde{\chi}^{\scriptscriptstyle 0}_1 $ and $\tilde{\chi}^{\scriptscriptstyle 0}_2 $-$\tilde{\chi}^{\scriptscriptstyle \pm}_1 $ mass differences (that are crucial quantities entering the phase-space factor of the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ widths) are shown versus $\tan \beta $, for the scenarios $\tilde{A}$-$\tilde{H}^+$ Some influence of $\tan \beta $ is also observed in the scalar mass spectrum (\hbox{cfr.}{}\ Appendix B). In \hbox{Figs.~}{}\ref{wbrvstgba}--\ref{wbrvstgbhp}, the behaviour of $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ widths and BR's as functions of $\tan \beta $ is shown. In the scenario $\tilde{A}$, we can distinguish three different regimes. For $\tan \beta \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$<$}\; 1.3$, the tree level decay into a light Higgs $h^{\scriptscriptstyle 0} $ is allowed and the corresponding BR is greater than 80\%. Around $\tan \beta = 1.5$, there is a small transition region where the charged-leptonic channel saturates the BR, since the decay into a $\tilde{e}_{\scriptscriptstyle L,R} $ is the only possible two-body real channel (this corresponds to the scenario A described above). For larger $\tan \beta $, there is a combined effect of the relative decreasing of the sneutrino mass with respect to the charged slepton masses (\hbox{cfr.}{}\ Appendix B) and the increasing Z-ino component in the $\tilde{\chi}^{\scriptscriptstyle 0}_1 $, that enhances the $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \nu_{\scriptscriptstyle \ell} \bar{\nu}_{\scriptscriptstyle \ell} $ decay. Therefore, for large $\tan \beta $ the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ gives rise mostly to missing energy and momentum. In the scenario $\tilde{B}$ (\hbox{Fig.~}{}\ref{wbrvstgbb}), $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 q\bar{q} $ is dominant in the whole $\tan \beta $ range considered, although the charged lepton channel has a considerable BR for $\tan \beta \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; 5 \div 10$. In scenario $\tilde{C}$ (\hbox{Fig.~}{}\ref{wbrvstgbc}), the hadronic channel is the main one for $\tan \beta \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$<$}\; 1.5$-2, while the leptonic channels get comparable to the former at higher $\tan \beta $. Note that, for $\tan \beta \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; 10$, also cascade decays into a $\tilde{\chi}^{\scriptscriptstyle \pm}_1 $ give a sizeable contribution. As for scenario $\tilde{D}$, we note in \hbox{Fig.~}{}\ref{wbrvstgbd} a maximum in both the invisible and the hadronic BR's curves corresponding to a deepening of the charged-leptonic one for $\tan \beta \simeq 10$. This is due to the sudden opening of the channel $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow e^{\scriptscriptstyle \pm} \tilde{e}^{\scriptscriptstyle \mp}_{\scriptscriptstyle L,R} $. In \hbox{Figs.~}{}\ref{wbrvstgbhm} and \ref{wbrvstgbhp}, we study the scenarios $\tilde{H}^{\mp}$. Here, we observe again a BR pattern closely connected to the $Z^{\scriptscriptstyle 0} $ BR's with some deviation due to the possible presence of a light chargino in the cascade decays (see also \hbox{Fig.~}{}\ref{massdiffvstgb}). In the scenario H$^{\rm -}$, the cascade decays contribute considerably at large $\tan \beta $, while, in the scenario H$^{\rm +}$, they decrease with $\tan \beta $. Note the strong (although not phenomenologically relevant) deepening of the radiative decay width at $\tan \beta \simeq 2.2$, due to destructive interference among various contributions. \section{Decreasing the Higgs masses} \label{sec:Higgs} In this section we study the sensitivity of the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ decay widths and BR's to a $m_{h^{\scriptscriptstyle 0}} $ change. In particular, we set $m_{A^{\scriptscriptstyle 0}} = M_{\scriptscriptstyle Z} $ which, compared to the case $m_{A^{\scriptscriptstyle 0}} = 3 M_{\scriptscriptstyle Z} $ studied in section 3, corresponds to a lowering of $m_{h^{\scriptscriptstyle 0}} $ down to 40-70 GeV, at $\tan \beta = 1.5$, and to 90-91 GeV, at $\tan \beta = 30$ in the considered range of $m_0$ and $M_2$ (\hbox{cfr.}{}\ section 3). In \hbox{Fig.~}{}\ref{brLH}a, we show the BR($\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} $), when $m_{A^{\scriptscriptstyle 0}} $ is lowered down to $M_{\scriptscriptstyle Z} $. The corresponding reduction of the threshold for the decay $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} $ considerably extends (with respect to \hbox{Fig.~}{}\ref{brnlh}) the area where BR$(\tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} ) > 30\%$ in the $(\mu, M_2)$ plane, down to regions of interest for LEP2 physics. In particular, this happens in the regions where the $\tilde{\chi}^{\scriptscriptstyle 0}_{1,2} $ gaugino components are large, which implies a considerable decrease in the BR's for all the other decay channels in these regions (as can be checked by comparing \hbox{Figs.~}{}\ref{brLH}b-d with \hbox{Figs.~}{}\ref{brnee}a, \ref{brnvv}a and \ref{brnqq}a). On the contrary, the situation is not altered in the higgsino region, where $m_{\tilde{\chi}^0_1} $ and $m_{\tilde{\chi}^0_2} $ tend to be degenerate, and the decay $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} $ is not allowed. In \hbox{Figs.~}{}\ref{wbrvsm0a_LH} and \ref{wbrvstgba_LH}, we study the influence of the $m_{A^{\scriptscriptstyle 0}} $ decrease in the scenarios $A/\tilde{A}$ considered in section 4. The scenario A is the only one, out of the six defined in \hbox{Tab.~}{}\ref{lep2scen}, that falls in the area of large BR($\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} )$ for $m_{A^{\scriptscriptstyle 0}} = M_{\scriptscriptstyle Z} $. In \hbox{Fig.~}{}\ref{wbrvsm0a_LH}, one can see the $m_0$ dependence of widths and BR's (already studied in \hbox{Fig.~}{}\ref{wbrvsm0a} for $m_{A^{\scriptscriptstyle 0}} = 3M_{\scriptscriptstyle Z} $), when one sets $m_{A^{\scriptscriptstyle 0}} = M_{\scriptscriptstyle Z} $. The influence of $m_0$ on the decay width into a Higgs comes from the radiative correction to $m_{h^{\scriptscriptstyle 0}} $ through the stop mass (\hbox{cfr.}{}\ Appendix B). After the low-$m_0$ range, where the invisible decay $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\nu}_{\scriptscriptstyle e,L} \nu_{\scriptscriptstyle e} $ is still dominant, we have an intermediate range $M_{\scriptscriptstyle Z} \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$<$}\; m_0 \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$<$}\; 350 {\rm\,GeV} $, where the $b\bar{b} +\not\!\! E $ signature (corresponding to the decay into $\tilde{\chi}^{\scriptscriptstyle 0}_1 h^{\scriptscriptstyle 0} $) is largely dominant. After that, the old pattern of \hbox{Fig.~}{}\ref{wbrvsm0a} is recovered with a large hadronic BR from $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 q\bar{q} $, through the $Z^{\scriptscriptstyle 0} $-exchange. As for the $\tan \beta $ dependence (\hbox{Fig.~}{}\ref{wbrvstgba_LH}), the main effect of lowering $m_{A^{\scriptscriptstyle 0}} $ with respect to \hbox{Fig.~}{}\ref{wbrvstgba} is an extension of the $\tan \beta $ range where the decay into a light Higgs is relevant, from about (1-1.4), up to about (1-2). Hence, the $\tan \beta $ range where the $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ decays in something visible is quite widened, in the scenario $\tilde{A}$. \section{The radiative $\protect\tilde{\chi}^{\scriptscriptstyle 0}_2$ decay} \label{sec:raddec} In this section, we study the BR for the decay $\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma$. Provided the mass difference $(m_{\tilde{\chi}^0_2} - m_{\tilde{\chi}^0_1} )$ is large enough as to give rise to a sufficiently energetic photon, this channel can produce a beautiful signature. A monochromatic photon plus missing energy and momentum should be observed. Although in all cases considered in the previous sections, BR($\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma$) never exceeds 15\%, it can reach values as large as 100\% in particular regions of the $\mu, M_2, \tan \beta $ space, as we are going to show. In \hbox{Fig.~}{}\ref{brnph}, the BR($\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma$) in the $(\mu, M_2)$ plane at fixed $m_0$ and $\tan \beta $ is studied. We will assume $m_{A^{\scriptscriptstyle 0}} = 3 M_{\scriptscriptstyle Z} $ everywhere. Indeed, although $m_{A^{\scriptscriptstyle 0}} $ sets the mass of the charged Higgs that flows in the virtual loops (\hbox{cfr.}{}\ \hbox{Fig.~}{}\ref{feyrad}), this parameter is the less critical in this study. We have checked that varying $m_{A^{\scriptscriptstyle 0}} $ in the range $(M_{\scriptscriptstyle Z} , 1 {\rm\,TeV} )$ can change BR($\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma$) by at most $\pm 10\%$ (with increasing BR when $m_{A^{\scriptscriptstyle 0}} $ grows). The scenarios studied in \hbox{Fig.~}{}\ref{brnph} assume either $\tan \beta = 1.5$ or 4. The BR for the radiative channel decreases substantially at larger $\tan \beta $. Furthermore, we set either $m_0 = M_{\scriptscriptstyle Z} $ or $3 M_{\scriptscriptstyle Z} $, as in the previous sections. One can distinguish a specific area where BR($\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma$) is large, that, particularly for large $m_0$ values (\hbox{cfr.}{}\ \hbox{Fig.~}{}\ref{brnph}c,d), evolves roughly around the $M_2 = - 2 \mu$ line. This corresponds to the region where $\tilde{\chi}^{\scriptscriptstyle 0}_1 $ is a pure higgsino-B, while $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ is mostly a photino (\hbox{cfr.}{}, \hbox{e.g.}{}, \hbox{Figs.~}{} 2.2 and 2.3 in \hbox{Ref.~}{}\cite{amb-mele}). This situation hinders all the tree-level decays, since the scalar-exchange decays require gaugino components in both $\tilde{\chi}^{\scriptscriptstyle 0}_1 $ and $\tilde{\chi}^{\scriptscriptstyle 0}_2 $, while the $Z^{\scriptscriptstyle 0} $-exchange ones need higgsino components. By the way, the different gaugino/higgsino nature of the two lightest neutralinos also depletes the $\tilde{\chi}^{\scriptscriptstyle 0}_1 \tilde{\chi}^{\scriptscriptstyle 0}_2 $ pair production in $e^{\scriptscriptstyle +}e^{\scriptscriptstyle -} $ collisions. In \hbox{Fig.~}{}\ref{brnph}, also note that a large fraction of the high-BR region lies in the area excluded by LEP1 searches. Nevertheless, one can single out particular situations (of interest for LEP2 physics and even beyond), where one can have very large BR's for the radiative channel. For instance, in \hbox{Fig.~}{}\ref{brvstgbfot}, we study two different scenarios versus $\tan \beta $: \\ (i) $\mu = -70 {\rm\,GeV} $, $M_2 = 130 {\rm\,GeV} $, $m_0 = M_{\scriptscriptstyle Z} , 3M_{\scriptscriptstyle Z} , 1 {\rm\,TeV} $; \\ (ii) $\mu = -120 {\rm\,GeV} $, $M_2 = 230 {\rm\,GeV} $, $m_0 = M_{\scriptscriptstyle Z} , 3M_{\scriptscriptstyle Z} , 1 {\rm\,TeV} $. \\ The scenario (i) is of interest for LEP2 searches. For instance, for $\tan \beta = 1.5$, one has $m_{\tilde{\chi}^0_1} \simeq 65 {\rm\,GeV} $ and $m_{\tilde{\chi}^0_2} \simeq 75 {\rm\,GeV} $. The scenario (ii) concerns heavier neutralino states: for $\tan \beta = 1.5$, one gets $m_{\tilde{\chi}^0_1} \simeq 113 {\rm\,GeV} $ and $m_{\tilde{\chi}^0_2} \simeq 127 {\rm\,GeV} $. Fig. \ref{brnph} shows that BR($\tilde{\chi}^{\scriptscriptstyle 0}_2 \rightarrow \tilde{\chi}^{\scriptscriptstyle 0}_1 \gamma$) is particularly large at moderate $\tan \beta $, although the total width can be as low as a few 10$^{-3}$ eV. The BR is enhanced by increasing $m_0$, since this makes all the other decays widths decrease further. On the other hand, one can check that, at $\tan \beta = 1$, $\tilde{\chi}^{\scriptscriptstyle 0}_1 $ and $\tilde{\chi}^{\scriptscriptstyle 0}_2 $ are almost degenerate, while their mass difference grows monotonically with $\tan \beta $. In order to have a sufficiently energetic photon, \hbox{e.g.}{}\ $E_{\gamma} \ge 10 {\rm\,GeV} $, one should restrict to $\tan \beta \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; 1.5$ in the case (i) and $\tan \beta \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; 1.35$ in the case (ii), which imply $(m_{\tilde{\chi}^0_2} - m_{\tilde{\chi}^0_1} ) \;\raisebox{-.5ex}{\rlap{$\sim$}} \raisebox{.5ex}{$>$}\; 10 {\rm\,GeV} $. A more in-depth study of the case of large radiative BR will be carried out in a forthcoming paper \cite{amb-mele2}. \acknowledgments We thank Guido Altarelli for constant encouragement and discussions.
\section{INTRODUCTION} The question of how fast a pulsar can spin has recently been discussed by several authors (see {\it e.g.\ } Friedman \& Ipser 1992, Cook et al. 1994a, b). Independently of the possible mechanisms of pulsar spin-up, for any particular equation of state there is always an upper limit for the final allowed rotational speed, $\Omega_{\rm max}$. The limit is determined either by the mass shedding condition, {\it i.e.} that the angular velocity of the configuration, stable with respect to axisymmetric perturbations, equals the Keplerian velocity at the surface, or by the condition of the onset of non-axisymmetric (e.g. gravitational radiation reaction) instabilities. In this paper we discuss the limit given by the mass shedding condition. Haensel and Zdunik (1989; hereafter HZ) noticed that for realistic equations of state of dense matter, the numerically calculated values of the shedding limit $\Omega_{\rm max}$ can be fitted, with an accuracy better than 5\%, by an empirical formula $$ \Omega_{\rm max} = {\cal C}_{\Omega}\left ({GM_{\rm max}}\over {R_{\rm max}^3}\right )^{1/2}, \eqno(1.1) $$ where $M_{\rm max}$ is the maximal mass of the non-rotating neutron stars with the same equation of state, $R_{\rm max}$ is the radius corresponding to $M_{\rm max}$, and ${\cal C}_{\Omega}$ is a dimensionless phenomenological constant, independent of the equation of state. $\Omega_{\rm max}$ is an angular velocity of rigid rotation as measured by a stationary observer at infinity. HZ determined that the best fit is for ${\cal C}_{\Omega} = 0.67$. The calculations of Friedman, Ipser and Parker (1989), performed for a very broad set of equations of state, yielded ${\cal C}_\Omega=0.66$, quoted in Friedman (1989), Friedman and Ipser (1992). It should be mentioned, that the value of ${\cal C}_\Omega=0.62$, quoted in the original paper of Friedman, Ipser and Parker (1989), was actually a very rough estimate of ${\cal C}_\Omega$ (J.L. Friedman, private communication). Calculations based on recent very accurate numerical methods are in good agreement with the original HZ choice of $C_\Omega$. For example, values of $\Omega_{\rm max}$ calculated for several equations of state by Lattimer et al. (1990) differ from the HZ version of the empirical formula by less than 4\%. Most recent calculations in the full framework of General Relativity by Salgado et al. (1994a,b; hereafter SBGH) based on the spectral methods (Bonazzola et al. 1993) show that configurations with causal equations of state (EOS) satisfy (1.1) with ${\cal C}_{\Omega} =0.67$ with an accuracy of better than 5\% (Haensel, Salgado and Bonazzola 1995, hereafter HSB). Results obtained for causal EOS by Cook et al. (1994b) lead to very similar ``best fit" value of ${\cal C}_\Omega$. HSB found that the empirical relation with ${\cal C}_\Omega=0.67$ fails for configurations constructed with non-causal equations of state, where the sound speed, $(\partial P/\partial \rho)^{1/2}_{ S}$, may be greater than the speed of light within the neutron star models. Empirical relations of universal character, valid for a broad range of realistic equations of state of dense matter, can obtained also for other bulk parameters of neutron star models. For example, a simple and surprisingly good universal relation, connecting the maximum moment of inertia for slow, rigid rotation, $I_{\rm max}$, to the mass and radius of a static configuration with maximum allowable mass (which is different from that with $I_{\rm max}$ !) was pointed out by Haensel (1992). The existence of such universal empirical relations is of practical importance: for any realistic EOS, empirical formulae enable rapid and still quite precise calculation of the upper bounds on global parameters of neutron stars from the easily calculated parameters of the static configuration with maximum allowable mass. A different type of a universal formula was obtained by Ravenhall and Pethick (1994). Their formula, valid for a broad range of realistic equations of state of dense matter, and useful for all except lightest neutron stars, expresses the moment of inertia in terms of stellar mass and radius. The few attempts to explain the empirical relation, Eq. (1.1), were not concluded with satisfactory results. The most elaborate discussion was presented by Weber and Glendenning (1991; 1992). They used numerical models of rotating neutron stars calculated in the slow rotation approximation of Hartle and Thorne (1968) to show that these also obey the empirical formula, albeit with ${\cal C}_{\Omega} \approx 0.75$. Although this is obviously an important result, still lacking is a clear physical understanding of the problem. In a recent work Glendenning and Weber (1994) derived a formula which relates $\Omega_{\rm max}$ to $M_{\rm max}/ R_{\rm max}^3 $ , in the slow rotation approximation, where $M_{\rm max}$ and $R_{\rm max}$ correspond to rotating configurations, but they do not discuss the connection between this relation and the formula (1.1). As we shall see below the slow rotation approximation cannot be applied consistently to maximally rotating neutron stars, as pointed out already by Hartle and Thorne (1968). \section{THE EMPIRICAL RELATIONS BETWEEN NON-ROTATING AND MAXIMALLY ROTATING CONFIGURATIONS} One can see that, to a good approximation, the ``constant" ${\cal C}_{\Omega}$ is in fact a slowly varying function of one variable, characterizing the EOS of dense matter, ${\cal C}_\Omega{\rm (EOS)}$ (Table 1 and HSB). Indeed Fig. 1 shows the values of ${\cal C}_{\Omega}$ as given by eq. (1.1) for 12 maximally rotating models calculated by SBGH, as a function of the parameter $$ x_{\rm s} = {2 G M_{\rm max}\over R_{\rm max} c^2} \eqno (2.1) $$ for static maximum mass configurations. It is clear that ${\cal C}_\Omega({\rm EOS})$ is slowly varying, and for the range of parameters of interest, can be well approximated by a monotonic function of $x_{\rm s}$. If one restricts to realistic EOS, which are both causal and stiff enough to support observed masses of pulsars, the range of relevant $x_{\rm s}$ becomes rather narrow and the approximation of ${\cal C}_{\Omega}({\rm EOS})$ by a constant is quite satisfactory (see HSB). It is well known that for realistic EOS, rotation increases the value of maximum mass of a neutron star by about 20$\%$. We find however that, with a very good approximation, there exists a {\it universal} relation between the maximum mass of non-rotating neutron star configuration and the mass of a configuration rotating with an angular speed $\Omega_{\rm max}$. The relation is: $$M_{\rm max}({\rm rot}) = {\cal C}_M M_{\rm max}({\rm stat}) , \eqno (2.2) $$ where $ M_{\rm max}({\rm rot})$ and $ M_{\rm max}({\rm stat})$ are respectively the mass of the configuration in maximum rotation and the maximum mass of the static neutron star for the same equation of state. Of course, a similar relation holds trivially for any specific EOS, and yields a specific value of the proportionality constant, ${\cal C}_M({\rm EOS})$. The values of ${\cal C}_M ({\rm EOS}) $, calculated for a broad set of EOS using the results of SBGH, are given in Table 1. The best fit of relation (2.2) to numerical results of SBGH is obtained for ${\cal C}_M=1.18$. Then, relation (2.2) reproduces exact results within better than 3\% (Fig.2). A similar relation is found for the radii of the corresponding configurations: $$R_{\rm max}({\rm rot}) = {\cal C}_R R_{\rm max}({\rm stat}) , \eqno (2.3) $$ where $R_{\rm max}({\rm rot})$ and $R_{\rm max}({\rm stat})$ are the radii of the maximally rotating and the static configurations respectively. The best fit value of a {\it constant} ${\cal C}_R$ in relation (2.3), based on numerical results of SBGH for causal EOS, is ${\cal C}_R=1.34$. Relation (2.3) reproduces then exact results for {\it causal} EOS within better than 4\% (Fig. 3). As in the case of relation (2.2), one can introduce the factor ${\cal C}_R ({\rm EOS})$ (Table 1). The dependence of ${\cal C}_R$ on the EOS can be well approximated by a monotonically decreasing function of $x_{\rm s}$ - this is visualized in Fig. 4. No such a trend was found for ${\cal C}_M{\rm (EOS)}$), Fig. 5. Our analysis have been based on a set of numerical results, obtained for twelve realistic EOS in SBGH. We have restricted to the SBGH set in order to keep the homogeneity of the sample of numerical results. Numerical results obtained by other authors (Friedman, Ipser \& Parker 1989, Cook et al. 1994b) are found to be slightly different from those of SBGH. This is due to the fact, that the precise determination of the maximally rotating configuration -- which requires very high precision of numerical procedure -- turns out to be sensitive to such details as, e.g., the interpolation method used for the determination of the maximum frequency model (Stergioulas \& Friedman 1995, E. Gourgoulhon, private communication). However, uncertainties resulting from this dependence on the specific sample of numerical results obtained for realistic EOS, are consistent with the precision of relations (2.2), (2.3), discussed above. One can easily see that the empirical ${\cal C}_\Omega$ constant may be obtained, within a very good approximation, from the formula: $$ {\cal C}_\Omega \simeq {\cal C}\equiv \left( { {\cal C}_M \over {{\cal C}_R}^3 } \right)^{1/2}. \eqno (2.4) $$ As can be seen from Table 1, for causal EOS Eq. (2.4) reproduces the actual values of ${\cal C}_\Omega({\rm EOS})$ within better than 5\%. In the case when one considers also non-causal EOS, the precision of approximation (2.4) worsens to 7\%. Relation (2.4) implies, that the angular velocity for mass shedding is approximated by the formula $$ \Omega_{\rm max} \simeq\left({G M_{\rm max}({\rm rot})\over R_{\rm max}^3({\rm rot})}\right)^{1/2}~. \eqno (2.5)$$ The value of $\Omega_{\rm max}$ can be thus well approximated by the frequency of a particle in stable circular orbit at the equator of a fictitious {\it non-rotating} star of mass equal to $M_{\rm max}({\rm rot})$ and of the radius equal to the equatorial radius of the maximally rotating configuration, $R_{\rm max}({\rm rot})$ (in the case of non-rotating star the general relativistic formula for the particle frequency is identical to the newtonian one, see {\it e.g.} Misner, Thorne \& Wheeler 1973). Up to now, we considered only {\it realistic} EOS of dense matter. It is instructive to study also the case of polytropic configurations with an equation of state in the form $P=K {n_{\rm b}}^{\Gamma}$, where $n_{\rm b}$ is baryon density of matter. Extensive calculations of rotating polytropic models were presented in Cook et al. (1992, 1994a). For a fixed value of $\Gamma$, static and rotating models exhibit useful scaling properties with respect to change of $K$ (Cook et al. 1992, 1994a). Consequently, the relativistic parameter $x_{\rm s}$, and ${\cal C}_M({\rm EOS})$, ${\cal C}_R({\rm EOS})$ turn out to be functions of $\Gamma$ only; they are given in Table 2. For a fixed value of $\Gamma$, relations (1.1), (2.2), and (2.3) are thus exact, with $\Gamma$-dependent values of the numerical coefficients. The realistic EOS are not polytropes, and their local adiabatic index, $\Gamma=(n_{\rm b}/P){\rm d}P/{\rm d}n_{\rm b}$, depends on the density, $\Gamma=\Gamma(n_{\rm b})$. Two specific examples of the density dependence of $\Gamma$ are shown in Fig. 6. In both cases, the value of $\Gamma$ varies within the relevant interval of $n_{\rm b}$ by more than $30\%$ of its maximum value. Clearly, such EOS cannot be represented by single polytropes. This explains, e.g., the lack of monotonic trend in the dependence of ${\cal C}_M({\rm EOS})$ on $x_{\rm s}({\rm EOS})$ for realistic EOS, displayed in Fig. 5. Such a non-monotonic, irregular behavior, characteristic of realistic EOS, is to be contrasted with a monotonic dependence ${\cal C}_M(x_{\rm s})$ for polytropes, Table 2. In spite of this irregular behavior of ${\cal C}_M({\rm EOS})$ for realistic EOS, the values of ${\cal C}({\rm EOS})$ for realistic EOS, calculated from (2.4), show a clear trend for a monotonic increase with $x_{\rm s}$. This results from a clear trend in ${\cal C}_R({\rm EOS})$ to decrease with $x_{\rm s}$, which -- magnified by the third power within the bracket of formula (2.4) -- dominates the ${\cal C}$ -- $x_{\rm s}$ relation. The trend for a monotonic increase of ${\cal C}_\Omega$ with $x_{\rm s}$, Fig. 1, can be well reproduced by a linear function ${\cal C}_\Omega^{\rm lin}(x_{\rm s})= 0.468 + 0.378x_{\rm s}$. This function, obtained by a least squares fit to the points shown in Fig. 1, reproduces ${\cal C}_\Omega({\rm EOS})$ (and $\Omega_{\rm max}({\rm EOS})$, if inserted in Eq. (1.1)) with precision better than $1.5\%$, with typical relative error being less than $1\%$. It should be stressed, however, that such a linear approximation ${\cal C}_\Omega^{\rm lin}(x_{\rm s})$ is valid only within a rather narrow interval of $x_{\rm s}$, characteristic of realistic EOS, $0.45<x_{\rm s}<0.7$. In contrast to the one-parameter empirical formula with ${\cal C}_\Omega=0.67$, ${\cal C}_\Omega^{\rm lin}(x_{\rm s})$ does not work for the EOS of the free neutron gas (see HSB). \section{THE SLOW ROTATION APPROXIMATION} Weber and Glendenning (1991, 1992) tried to derive the empirical formula, equation (1.1), by using the slow rotation approximation of the Einstein equations. One should notice however that a neutron star rotating with $\Omega_{\rm max}$ may not be really considered to be a slow rotator (Hartle \& Thorne 1978). Indeed, let us define a dimensionless angular velocity $\Omega_*$, $$ \Omega_*^{~2} \equiv {\Omega}^2/\left({GM\over R^3}\right), \eqno (3.1)$$ where $M$ is the mass of the static star, $R$ its radius and ${\Omega}$ is the angular velocity measured by observers at infinity. The assumption of the slow rotation means that (Hartle 1967), $$ \Omega_*^{~2} \ll 1. \eqno (3.2)$$ From equation (1.1) with ${\cal C}_{\Omega}=0.67$ it follows that $$ \Omega_{*\ \rm max}^{~2}\approx 0.45, \eqno(3.3) $$ so that the use of slow-rotation approximation for maximally rotating configurations requires some additional justification. It is unlikely that this approximation will give accurate results considering the fact that at the stellar surface the linear speed of rotation is a significant fraction of the speed of light (Hartle \& Thorne 1968): $$ {v_S \over c} = {1\over \sqrt 2} \left({x_s \over 1 - x_s}\right)^{1/2} \eqno(3.4) $$ For the maximally rotating models of SBGH one gets typically $v_S/c \approx 0.7$. In the slow rotation approximation, one can obtain for $\Omega_{\rm max}$ an expression of the form, which seems to be similar to that of the empirical formula (1.1) (see also Glendenning and Weber 1994): $$ \Omega_{\rm max} = {\cal C}_{\rm sr}\left ({GM_{\rm max}({\rm rot}) \over R_{\rm max}^{~3}({\rm rot})}\right)^{1/2} + {\cal O}(\Omega_*^3), \eqno (3.5)$$ with $$ {\cal C}_{\rm sr} \approx \left [ 1 + {{I}\over {MR^2}} {{R_G}\over R}\left(1 - 2.5{I\over MR^2} \left({R\over R_G}\right)^4 \left(1 -{R_G\over R}\right)Q^{'2}_2(u)\cdot \left(1-{{QMc^2}\over J^{2}}\right) \right) \right]^{-{1\over 2}}~. \eqno (3.6)$$ Here $R_G$ is the gravitational radius, $$ R_G = {{2GM}\over c^2}, \eqno (3.7)$$ $J$ is the angular momentum, $I$ the moment of inertia, $Q$ is the quadrupole moment of the rotating configuration, and $Q^{'2}_{2}(u)$ is the derivative, with respect to $u=1-R/R_G$, of the associated Legendre function of the second kind. The quantities $M$ and $R$ are those for the maximally rotating configuration, but -- within our approximation --- we can as well replace them by $M_{\rm max}({\rm stat})$, $R_{\rm max}({\rm stat})$. Equations (3.5) and (3.6) give $\Omega_{\rm max}$ in terms of the mass, equatorial radius, moment of inertia, angular momentum and the quadrupole moment of the maximally {\it rotating} configuration. At first glance (and neglecting terms $\sim \Omega_*^{~3}$ and higher), expression (3.5 -- 3.6) may seem to be similar to empirical formula (1.1). It has been shown by Abramowicz and Wagoner (1978) and recently confirmed by Ravenhall and Pethick (1994) that moments of inertia, expressed in the units of $MR^2$, are - to a good approximation - functions of only $x=R/R_G$. Moreover, results of SBGH show that ${QMc^2}/ J^{2}$ is a decreasing function of $x_s$. Expression (3.5) seems thus to possess similar properties as the empirical formula. Unfortunately {\it it is not} the empirical formula since it involves $M_{\rm max}$ and $R_{\rm max}$ of {\it rotating} configurations, and therefore it corresponds to some approximation of eq. (2.5) and not to eq. (1.1). Further expansions of $M_{\rm max}({\rm rot})$ and $R_{\rm max}({\rm rot})$ in $\Omega_*$ are not justified - in view of the large value of this parameter (notice that rotation increases the value of $R_{\rm max}$ by some 30\%). So, only some external ``empirical input" (such as assumption of a ``typical" effect of rotation on $M_{\rm max}$, $R_{\rm max}$, and on eccentricity of rotating star, made in Weber \& Glendenning (1992)) can lead to an expression of type (1.1). A consistent application of the slow rotation approximation cannot reproduce the empirical formula for $\Omega_{\rm max}$. This is not very surprising if one considers that one should get to the 8-th order in $\Omega_*$ to get a precision of 4\%, characteristic of empirical formula. \section{DISCUSSION AND CONCLUSIONS} The proportionality constant appearing in the empirical formula for $\Omega_{\rm max}$ for realistic EOS of dense matter, is in fact a function of the relativistic parameter $x_s$. In the range of parameters describing maximally rotating configurations with causal equations of state ${\cal C}_{\Omega}$ is a rather slowly varying function of $x_s$, which results in a high precision of empirical formula with an appropriate choice of a universal proportionality constant. We found universal relations connecting maximal masses and corresponding radii of static neutron stars, calculated for realistic EOS of dense matter, with those of maximally rotating configurations. The empirical formula follows from those relations. Although the slow rotation approximation allows one to reproduce some of the properties of the empirical formula this approximation is not appropriate for maximally rotating configurations. It is a great pleasure to thank Silvano Bonazzola for his role in stimulating this research. We are grateful to him, Eric Gourgoulhon, Marcelo Salgado and R\'emi Hakim for very useful discussions. We thank the second referee of this paper for useful suggestions and comments. This research was partially supported by the Polish State Committee for Scientific Research (KBN) grant, and by PICS/CNRS no. 198 ``Astronomie Pologne".
\section{Introduction} In recent years, various kinds of nonequilibrium lattice models exhibiting a continuous phase transition from a reactive phase into an absorbing (inactive) phase have been studied extensively \cite{Liggett,Dickb}. Most of models investigated are found to belong to the directed percolation (DP) universality class \cite{Jan,Grass82,Cardy,Grin}. A common feature of these models is that the absorbing phase consists of a {\em single} absorbing state. Even some models with infinitely many absorbing states also exhibit critical behavior in the DP universality class \cite{Jen932,Jen933}. Only a few models have been studied which are not in the DP universality class. Those are the model A and B of probabilistic cellular automata \cite{Grass84,Grass89}, nonequilibrium kinetic Ising models with two different dynamics \cite{Meny94,Meny95}, and interacting monomer-dimer models \cite{Park94,Park951,Park952}. Numerical investigations show that they belong to the same but non-DP universality class. These models share a common property that the absorbing phase consists of two {\em equivalent} absorbing states. Recently, the branching annihilating random walks (BAW) have been studied intensively \cite{Bram,Sud,Taka,Jen934,Jen931,Jen941}. The BAW model is a lattice model where a walker can hop to a nearest neighbor site with probability $p$ and branch with $n$ offsprings in nearest neighborhood with probability $1-p$. Two walkers annihilate immediately when they meet. In general, this model exhibits a continuous phase transition from an active state into the absorbing state (vacuum) at finite hopping probability $p_c$. Even though the BAW model has a single absorbing state, its critical property depends on the parity of the number of offsprings created in a branching process. Dynamics of the BAW models with even $n$ conserve the number of walkers modulo 2, while the BAW models with odd $n$ evolve without any conservation. The BAW models with odd $n$ exhibit the DP critical behavior \cite{Taka,Jen934}, while the BAW models with even $n$ exhibit the same non-DP behavior as in the models with two equivalent absorbing states \cite{Grass89,Park94,Park951,Park952,Jen941}. One can find that the total number of kinks are conserved modulo 2 in models with two equivalent absorbing states. Therefore this conservation law may be responsible for the non-DP critical behavior. The critical exponents characterizing the non-DP behavior are measured accurately by extensive Monte Carlo simulations for the BAW model with $n=4$ in one dimension \cite{Jen941}. Unfortunately, the BAW model with $n=2$ which is simpler does not have an active phase ($p_c=0$) \cite{Sud,Taka2}. Recently, ben-Avraham {\em et al} \cite{Redner} introduce another parameter which controls the two-walker annihilation process and show that the BAW model with $n=2$ exhibits a continuous phase transition at a finite value of $p_c$ except for the case of infinite annihilation rate (the ordinary BAW model). But the values of the critical exponents are not reported there. As the annihilation rate becomes smaller, it is clear that the system gets more active (more walkers survive) so the critical probability $p_c$ goes higher. In the ordinary BAW models, there are two competing elementary processes, i.e.~hopping and branching. The hopping process does not increase the number of walkers but can decrease it by two-walker annihilations. As the hopping probability becomes larger, the system tends to trap into the absorbing state (vacuum). The branching process may increase the number of walkers so it can make the system more active. That is why the BAW model with one offspring exhibits a phase transition from an active state into vacuum as $p$ becomes larger. Therefore one may argue that increasing the number of offsprings in the branching process will make the system more active so the critical probability $p_c$ will go higher as $n$ becomes bigger. However, numerical and analytical study of the BAW models with $n$ offsprings show that $p_c$ is not a monotonically increasing function of $n$. In fact, the values of $p_c$ are 0.1070(5), 0, 0.459(1), 0.7215(5), and 0.718(1) for $n=1-5$ \cite{Taka,Jen934,Jen941}. These results contradict the conventional wisdom mentioned above and there is no consistent manner in making the system more active by changing the number of offsprings. In order to understand this rather surprising result, we introduce a model which interpolates the BAW models with one offspring and with two offsprings. In this model, the branching process creates one offspring or two offsprings with relative ratio. We map out the phase diagram by dynamic Monte Carlo simulations which shows a reentrant phase transition from vacuum to an active state and finally into vacuum again as the relative rate of the two-offspring process increases at fixed hopping probability. The second phase transition occurs at quite high rates of the two-offspring process (more than $80\%$). We argue that the second transition is due to the static reflection symmetry of the two-offspring branching process (one offspring to the left and the other to the right of the branching walker: {\em static branching}). This reentrant second transition disappears and our conventional wisdom is recovered when the dynamic reflection symmetry is introduced (both offsprings to the left or to the right of the branching walker with equal probability: {\em dynamic branching}). In the next section, we describe the BAW model with one and two offsprings with relative ratio. Dynamic Monte Carlo results are discussed which show the reentrant phase diagram. In Sec.~III, the BAW model with dynamic branching is introduced. Our numerical results for $n=2$ find the existence of the non-DP critical behavior at finite hopping probability. In Sec.~IV, we study the BAW model with one and two offsprings created by dynamic branching. The reentrant behavior disappears entirely as expected. We conclude in Sec.~V with brief summary. \section{The BAW model with one and two offsprings: Static branching} We consider the BAW model with one and two offsprings with relative ratio in one dimension. The evolution rules of this model are given as follows. First, choose a walker at random. It may hop to a randomly chosen nearest neighbor site with probability $p$. If this site is already occupied by another walker, both walkers annihilate immediately. With probability $1-p$, the randomly chosen walker creates two offsprings symmetrically at nearest neighbor sites with relative probability $r$ ($0\le r \le 1$) or creates one offspring at a randomly chosen nearest neighbor site with relative probability $1-r$. When an offspring is created on a site already occupied, both walkers annihilate immediately. The case $r=0$ corresponds to the BAW model with one offspring which exhibits a continuous phase transition from an active phase into vacuum at $p\simeq 0.1070$ \cite{Jen934}. The other limiting case $r=1$ corresponds to the BAW model with two offsprings which does not have an active state at finite values of $p$ \cite{Sud,Taka2}. We perform dynamic Monte Carlo simulations for this model with various values of $r=0$, 0.25, 0.50, 0.75, 0.8, 0.85, 0.9, 0.95, and 1. We start with two nearest neighbor walkers at the central sites of a lattice. Then the system evolves along the dynamical rules of the model. After one attempt of hopping or branching on the average per lattice site (one Monte Carlo step), the time is incremented by one unit. A numer of independent runs, typically $10^5$, are made up to 2000 time steps for various values of $p$ near the critical probability $p_c$. Most runs, however, stop earlier because the system gets into the absorbing state. We measure the survival probability $P(t)$ (the probability that the system is still active at time $t$), the number of walkers $N(t)$ averaged over all runs, and the mean-square distance of spreading $R^2 (t)$ averaged over the surviving runs. At criticality, the values of these quantities scale algebraically in the long time limit \cite{Grass79} \begin{eqnarray} P(t) &\sim& t^{-\delta},\nonumber \\ N(t) &\sim& t^{\eta},\\ R^2(t) &\sim& t^{z},\nonumber \end{eqnarray} and double-logarithmic plots of these values against time show straight lines. Off criticality, these plots show some curvatures. More precise estimates for the scaling exponents can be obtained by examining the local slopes of the curves. The effective exponent $\delta(t)$ is defined as \begin{equation} -\delta(t) = \frac{\log \left[ P(t) / P(t/b) \right]}{\log ~b} \end{equation} and similarly for $\eta (t)$ and $z(t)$. In Fig.~1, we plot the effective exponents against $1/t$ with $b =5$ for $r=0.25$. Off criticality these plots show upward or downward curvatures. From Fig.~1, we estimate $p_c = 0.1265(5)$ and dynamic exponents $ \delta = 0.158(2)$, $\eta = 0.310(3)$, $z = 1.27(1)$. These values are in an excellent accord with those of the DP universality class; $\delta=0.1596(4), \eta=0.3137(10)$, and $z=1.2660(14)$ (see reference \cite{Jen941}). This result supports the conjecture that models with a single absorbing state and no conservation laws should belong to the DP universality class. Similarly we determine the values of the critical probability $p_c$ and the dynamic exponents for various values of $r$. As expected from the above conjecture, the values of the dynamic exponents stay almost unchanged except for $r=1$. Estimates of $p_c$ for various values of $r$ are listed in Table I. The value of $p_c$ slowly increases as $r$ varies upto $\sim 0.75$ and drastically decreases to zero as $r$ approaches the value of unity. The $r-p$ phase diagram is drawn in Fig.~2 where a reentrant phase transition is explicitly shown. At fixed values of $p=0.11\sim 0.14$, the system undergoes phase transitions from vacuum to an active state and finally into vacuum again as $r$ becomes larger (more offsprings are created). The second transition occurs at quite high rates of the two-offspring branching process. This result implies that our conventional wisdom does not apply near $r=1$. The BAW model with two-offsprings can be solved exactly at $p=0$, where pairs of nearest neighbor walkers diffuse like the ordinary random walks, i.e.~$\cdots 001100\cdots \rightarrow \cdots 000110\cdots$ where `$1$' represents an occupied site and `$0$' a vacant site. When two pairs collide each other, they just bounce back. So the number of walkers is bounded in this model, in contrast to the BAW model with one offspring where the number of walkers is not bounded. So our conventional wisdom that the BAW model with more offsprings may be more active does not work. For $p>0$, these pairs annihilate by hopping processes so the system becomes inactive in the long time limit. We notice that picture of diffusing pairs is mainly due to the static reflection symmetry of the branching process. If the dynamic reflection symmetry is adopted instead of the static one, this diffusing-pair picture is no longer valid. In this case, a walker branches two offsprings both to the left or to the right of itself with equal probability. The reflection symmetry is not broken on average, but this branching process allows a formation of long chains of walkers, i.e.~$\cdots 00011000 \cdots \rightarrow \cdots 01111000\cdots$. So the number of walkers is not bounded from above and an active phase may appear at finite hopping probability. \section{The BAW model with two offsprings: Dynamic branching} We study the BAW model with two offsprings created by dynamic branching. The hopping process is the same as in the ordinary BAW models, but in the branching process a randomly chosen walker creates two offsprings both on the sites to the left or to the right of the walker with equal probability. We perform dynamic Monte Carlo simulations, starting from a pair of walkers at the central sites of a lattice. $10^6$ independent runs are made during 2000 time steps and we measure $P(t)$, $N(t)$, and $R^2 (t)$. In Fig.~3, we plot the effective exponents $\delta(t)$, $\eta(t)$, and $z(t)$ against $1/t$ with $b=5$. These plots clearly show the existence of an active phase. We estimate $p_c=0.5105(7)$ and the dynamic exponents $\delta=0.287(1)$, $\eta=0.000(3)$, and $z=1.155(5)$. These values are in an excellent accord with those of the non-DP universality class; $\delta=0.285(2)$, $\eta=0.000(1)$, and $z=1.141(2)$ for the BAW model with four offsprings. This result supports the conjecture that models with particle number conservation of modulo 2 should belong to the same but non-DP universality class. We also perform dynamic simulations with a different initial configuration. We start with a single walker at the center of a lattice. Conservation of the number of walkers of modulo 2 prevents the system from entering the absorbing state (vacuum). So the survival probability exponent $\delta$ must be zero. Our numerical results conclude that the dynamic exponents $\eta=0.283(5)$ and $z=1.15(1)$. These values also agree very well with those of the BAW model with four offsprings ($\eta=0.282(4)$, $z=1.140(5)$). Compared with the ordinary BAW model with two offsprings, it is clear that the static reflection symmetry is responsible for the nonexistence of an active phase and the dynamic reflection symmetry makes the system more active. To support this idea, we employ the mean field theory on spreading of the active region. Difference between these two models lies on the branching process. So we only consider the effect of different branching mechanism on the boundary of the active region. The boundary can move by branching of walkers at the boundary or nearby. For example, a branching process of a walker at the boundary increases the active region by one unit in the BAW model with the static symmetry. For the BAW model with the dynamic symmetry, the same process increases the active region by the same amount (one unit on average). However, a branching process of a walker near the boundary inside of the active region decreases the active region differently for these two models. Considering all possible cases that the boundary can move and using the mean field theory to assign a proper probability to each case, we find the outward velocity of the boundary as \begin{eqnarray} v_s &=& (1-p)\rho (1-\rho), \nonumber\\ v_d &=& (1-p)\rho (1-\frac{\rho^2}{2}), \end{eqnarray} where $v_s$ ($v_d$) is the outward velocity of the boundary for the model with the static (dynamic) symmetry and $\rho$ is the walker density. The first term comes from the branching process of a walker at the boundary and the second term near the boundary. The effect of hopping on the boundary velocity is omitted. We find that $v_s$ is always smaller than $v_d$. This result implies that the active region grows faster in the model with the dynamic symmetry, so this model should be more active than the model with the static symmetry. This mean field argument can be easily generalized to the models with general $n$ offsprings. As the critical probability depends enormously on the symmetry of branching processes (especially $n=2$), it is meaningless to ask whether $p_c$ monotonically increases with $n$ for the ordinary BAW models. These models have not been classified by the symmetry of branching processes. The branching process of the ordinary BAW model with one offspring basically belongs to the process with the dynamic symmetry. If we compare the critical probabilities of the BAW models with dynamic branching only, we expect that $p_c$ will monotonically increase with $n$, i.e.~the dynamic branching process always makes the system more active. In the next section, we study the BAW model with one and two offsprings created by dynamic branching and examine whether the reentrant behavior seen in the case of static branching (Sec.~II) disappears. \section{The BAW model with one and two offsprings: Dynamic branching} We consider the BAW model with one and two offsprings branched dynamically with relative ratio. The evolution rules of this model are exactly the same as in Sec.~2 except that static branching is replaced with dynamic branching for the two-offspring branching process. We perform dynamic Monte Carlo simulations for $r=0.25$, 0.50, 0.75. Estimated values of $p_c$ are listed in Table II. As expected, $p_c$ monotonically increases as $r$ becomes larger (more offsprings are created). The reentrant phase transition disappears entirely in this model with dynamic branching (see Fig.~4). Of course, the critical behavior at the absorbing transitions belongs to the DP universality class for $0\le r <1$. \section{Summary} We study the BAW models with static (ordinary) branching and dynamic branching. With the static branching, the BAW model with one and two offsprings shows a reentrant phase transition from vacuum to an active state and finally into vacuum again as the relative rate of the two offspring process increases. We argue that this reentrant property originates from the static reflection symmetry of the two-offspring branching process. The ordinary BAW model with two offsprings does not have an active phase at finite values of hopping probability. We introduce the BAW model with two offsprings created by dynamic branching and show that this model exhibits a continuous phase transition from an active phase into vacuum at finite hopping probability. Its critical behavior belongs to the same non-DP universality class as in the ordinary BAW model with four offsprings. We also study the BAW model with one and two offsprings created by dynamic branching. The reentrant behavior disappears and our conventional wisdom is recovered as expected. Our results shed light on how the system can be active by different branching processes and the BAW model with dynamic branching may serve as another simple model exhibiting the non-DP critical behavior. \section*{acknowledgements} We wish to thank Heungwon Park for interesting discussions. This work is supported in part by the BSRI, Ministry of Education (Grant No.~95-2409) and by an Inha University research grant (1995).
\section{ Introduction} The purpose of this article is to describe a numerically stable Quantum Monte Carlo (QMC) algorithm to calculate zero-temperature imaginary time displaced Green functions: \begin{eqnarray*} G_{x,y}(\tau) = \Theta(\tau) \frac{ \langle \Psi_0 | c_{x}(\tau) c_{y}^{\dagger} | \Psi_0 \rangle} {\langle \Psi_0 | \Psi_0 \rangle} - \Theta(-\tau) \frac{ \langle \Psi_0 | c_{y}^{\dagger}(-\tau) c_{x} | \Psi_0 \rangle} {\langle \Psi_0 | \Psi_0 \rangle }, \end{eqnarray*} \begin{equation} {\rm where } \; \; \; \; c_{x}(\tau) = e^{ \tau (H - \mu N) } c_{x} e^{-\tau (H - \mu N)}. \end{equation} Here $ |\Psi_0 \rangle $ denotes the ground state of the considered Hamiltonian $H$, $c_{x}^{\dagger}$ creates an electron with quantum numbers $x$, $\Theta(\tau)$ is the Heaviside function and $\mu $ is the chemical potential which has to satisfy: \begin{equation} \label{chem} \lim_{\beta \rightarrow \infty} \frac{ {\rm Tr} \left( e^{-\beta (H - \mu N)} N \right) } { {\rm Tr} \left( e^{-\beta (H - \mu N)} \right) } = \frac{ \langle \Psi_0 | N | \Psi_0 \rangle } {\langle \Psi_0 | \Psi_0 \rangle }, \end{equation} in the zero-temperature limit $T \equiv 1/\beta \rightarrow 0$. The above equation implies that the chemical potential corresponding to the desired particle density, $n(\mu)$, has to be known prior to the simulation. In a metallic state $n(\mu)$ is in general not known a priori. However, in an insulating state at zero-temperature $n(\mu)$ is constant and in general known for chemical potentials within the charge gap $\Delta_c$. In this situation, the here described algorithm proves to be a powerful tool. The above $T=0$ Green functions have already been calculated with QMC methods by Deisz et al \cite{John}. Since their algorithm does not incorporate a numerical stabilization scheme, they are restricted to relatively small values of $\tau$ (i.e. $\tau = 2.5 $ in units of the hopping matrix element for the one-dimensional Hubbard model). This articles follows the work of Deisz et al. and describes a numerical stabilization scheme which allows one to calculate $T=0$ Green functions for arbitrary values of $\tau$. To demonstrate the efficiency of the algorithm, we calculate $G_{x,y}(\tau) $ for the two-dimensional half-filled Hubbard model: \begin{equation} \label{Hubb} H = \sum_{\vec{i},\vec{j}, \sigma} c_{\vec{i},\sigma}^{\dagger}T_{\vec{i},\vec{j}} c_{\vec{j},\sigma} + U \sum_{\vec{i}} \left( n_{\vec{i}, \uparrow} - \frac{1}{2} \right) \left( n_{\vec{i}, \downarrow} - \frac{1}{2} \right). \end{equation} The quantum numbers $\vec{i}$ and $\sigma$ denote lattice site and {\it z}-component of spin respectively, $n_{\vec{i},\sigma } = c_{\vec{i},\sigma}^{\dagger} c_{\vec{i}\sigma}$ , and $T_{\vec{i},\vec{j}} = -t$ if $\vec{i}$ and $\vec{j}$ are nearest-neighbors. In this notation half-band filling corresponds to $\mu = 0$. As a non-trivial test of the algorithm, one may fit the tail of $G_{x,x}(\tau)$ ($x = (\vec{i}, \sigma))$ to the form $ e^{-\tau \Delta_c}$ to obtain the charge gap $\Delta_c$. At $U/t = 4$ and after extrapolation to the thermodynamic limit, our QMC data yields $\Delta_c/t = 0.67 \pm 0.02 $. This value stands in good agreement with previously determined values of $\Delta_c$ \cite{Furukawa}. The article is organized in the following way. In the next section, we briefly describe the zero-temperature auxiliary-field QMC algorithm for the Hubbard model \cite{Koonin,Sandro,Imada,Assaad}. We then present our solution for the numerical stabilization of the time displaced Green functions. In section 3 we describe our calculation of the charge gap for the two-dimensional Hubbard model at $U/t = 4$. In the last section, we draw some conclusions and discuss the potential applications of the algorithm. \section{The zero-temperature QMC algorithm } Since the Hubbard model conserves particle number, and we are working in the canonical ensemble, one may factorize the chemical potential to obtain: \begin{equation} \label{factor} G_{x,y}(\tau) = \Theta(\tau) \left. \frac{ \langle \Psi_0 | c_{x}(\tau) c_{y}^{\dagger} | \Psi_0 \rangle } {\langle \Psi_0 | \Psi_0 \rangle } \right|_{\mu = 0} e^{\tau \mu} - \Theta(-\tau) \left. \frac{ \langle \Psi_0 | c_{y}^{\dagger}(-\tau) c_{x} | \Psi_0 \rangle } {\langle \Psi_0 | \Psi_0 \rangle }\right|_{\mu = 0} e^{\tau \mu}. \end{equation} Due to the above relation we consider the calculation of the $T=0$ Green functions at $\mu = 0$. The idea behind the zero temperature QMC algorithm is to filter out the ground state from a trial wave function $| \Psi_T\rangle $ which is required to be non-orthogonal to the ground state: \begin{equation} \frac{\langle \Psi_0 | c_{x}(\tau) c_{y}^{\dagger} | \Psi_0 \rangle } {\langle \Psi_0 | \Psi_0 \rangle } = \lim_{ \Theta \rightarrow \infty } \frac{ \langle \Psi_T |e^{-\Theta H } c_{x}(\tau) c_{y}^{\dagger} e^{-\Theta H } | \Psi_T \rangle } { \langle \Psi_T |e^{-2\Theta H } | \Psi_T \rangle }, \; \; \tau > 0. \end{equation} The QMC calculation of \begin{equation} G^{>}_{x,y}(\Theta,\tau) \equiv \frac{ \langle \Psi_T |e^{-\Theta H } c_{x}(\tau) c_{y}^{\dagger} e^{-\Theta H } | \Psi_T \rangle } { \langle \Psi_T |e^{-2\Theta H } | \Psi_T \rangle } \end{equation} proceeds in the following way. The first step is to carry out a Trotter decomposition of the imaginary time propagation: \begin{equation} \label{Trotter} e^{-2\Theta H } = \left( e^{- \Delta \tau H_t/2 } e^{- \Delta \tau H_U} e^{- \Delta \tau H_t/2 } \right)^{m} + O( (\Delta \tau)^2). \end{equation} Here, $H_t$ ($H_U$) denotes the kinetic (potential) term of the Hubbard model and $m \Delta \tau = 2\Theta$. Having isolated the two-body interaction term, $H_U$, one may carry out a discrete Hubbard Stratonovitch (HS) transformation \cite{Hirsch} to obtain: \begin{equation} \label{HS} e^{-\Delta \tau H_U} = C \sum_{\vec{s}} \exp \left( \sum_{x,y} c_{x}^{\dagger} D_{x,y}(\vec{s}) c_{y} \right), \end{equation} where $\vec{s}$ denotes a vector of HS Ising fields. For the Hubbard model (\ref{Hubb}), we take: \begin{equation} D_{\vec{i}\sigma, \vec{j}\sigma'} (\vec{s}) = \delta_{\sigma,\sigma'} \delta_{\vec{i},\vec{j}} {\rm cosh}^{-1} (\Delta \tau U/2) s_{\vec{i}} \sigma. \end{equation} The constant $C = {\rm exp} ( -\Delta \tau N U/2)/ 2^N $ for the $N$-site system will be dropped below. The imaginary time propagation may now be written as: \begin{eqnarray} & & e^{-2\Theta H } = \sum_{\vec{s}} U_{\vec{s}}(2\Theta, 0) + O( (\Delta \tau)^2) \nonumber \\ {\rm where} \; \; \; & &U_{\vec{s}}(2\Theta, 0) = \prod_{n = 1}^{m} e^{- \Delta \tau H_t/2 } \exp \left( \sum_{x,y} c_{x}^{\dagger} D_{x,y}(\vec{s}_n) c_{y} \right) e^{- \Delta \tau H_t/2 }. \end{eqnarray} In the above notation, $ G^{>}_{x,y}(\Theta,\tau)$ is given by: \begin{equation} G^{>}_{x,y}(\Theta,\tau) = \frac { \sum_{\vec{s}} \langle \Psi_T | U_{\vec{s}}(2\Theta, \Theta + \tau) c_x U_{\vec{s}}(\Theta + \tau,\Theta) c_y^{\dagger} U_{\vec{s}}(\Theta,0) | \Psi_T \rangle } {\sum_{\vec{s}} \langle \Psi_T | U_{\vec{s}}(2\Theta, 0) | \Psi_T \rangle } + O( (\Delta \tau)^2). \end{equation} The trial wave function is required to be a Slater determinant: \begin{equation} \label{Trial} |\Psi_T \rangle = \prod_{n=1}^{N_p} \left( \sum_x c_{x}^{\dagger} P_{x,n} \right) |0\rangle . \end{equation} Here $N_p$ denotes the number of particles and $P $ is an $N_s \times N_p$ rectangular matrix where $N_s$ is the number of single particle states. Since $U_{\vec{s}}(2\Theta, 0)$ describes the propagation of non-interacting electrons in an external HS field, one may integrate out the fermionic degrees of freedom to obtain: \begin{equation} \label{G>} G^{>}_{x,y}(\Theta,\tau) = \sum_{\vec{s}} P_{\vec{s}} \left[ \left( B_{\vec{s}} (\Theta + \tau, \Theta)\right) G_{\vec{s}}(\Theta,\Theta) \right]_{x,y} + O( (\Delta \tau)^2). \end{equation} In the above equation, \begin{eqnarray*} B_{\vec{s}} (\Theta_2, \Theta_1) = \prod_{n = n_1 + 1}^{n_2} e^{-\Delta \tau T /2} e^{D(\vec{s_n})} e^{-\Delta \tau T /2} \; \; {\rm where } \; \; n_1 \Delta \tau = \Theta_1 \; \; {\rm and} \; \; n_2 \Delta \tau = \Theta_2, \end{eqnarray*} \begin{eqnarray*} M_{\vec{s}} = P^{T} B_{\vec{s}} (2\Theta, 0 ) P, \end{eqnarray*} \begin{eqnarray*} P_{\vec{s}} = \frac{ \det(M_{\vec{s}}) } { \sum_{\vec{s}} \det(M_{\vec{s}}) } \end{eqnarray*} \begin{eqnarray*} {\rm and} \; \; \; \left( G_{\vec{s}}(\Theta,\Theta) \right)_{x,y} = \frac { \langle \Psi_T | U_{\vec{s}}(2\Theta, \Theta) c_x c_y^{\dagger} U_{\vec{s}}(\Theta,0) | \Psi_T \rangle } { \langle \Psi_T | U_{\vec{s}}(2\Theta, 0) | \Psi_T \rangle } = \\ \left( I - B_{\vec{s}} (\Theta,0)P M_{\vec{s}}^{-1} P^{T} B_{\vec{s}} (2\Theta,\Theta) \right)_{x,y}. \end{eqnarray*} Here $I$ is the unit matrix, $I_{x,y} = \delta_{x,y}$. In the same notation one obtains: \begin{eqnarray} \label{G<} G^{<}_{x,y}(\Theta,\tau) & \equiv & - \frac{ \langle \Psi_T |e^{-\Theta H } c_{y}^{\dagger}(\tau) c_{x} e^{-\Theta H } | \Psi_T \rangle } { \langle \Psi_T |e^{-2\Theta H } | \Psi_T \rangle } \nonumber \\ & = & \sum_{\vec{s}} P_{\vec{s}} \left[ \left( G_{\vec{s}}(\Theta,\Theta) - I \right) B_{\vec{s}}^{-1}(\Theta + \tau, \Theta) \right]_{x,y}, \; \; \tau > 0. \end{eqnarray} Summarizing, the zero-temperature imaginary-time Green function may be calculated from: \begin{equation} G_{x,y}(\tau) = \lim_{\Theta \rightarrow \infty } \left( \Theta(\tau) G^{>}_{x,y}(\Theta,\tau) + \Theta(-\tau) G^{<}_{x,y}(\Theta,-\tau) \right) + O( (\Delta \tau)^2). \end{equation} At half-band filling and due to particle hole symmetry, one may chose a trial wave function such that $P_{\vec{s}}$ is positive definite. $P_{\vec{s}} $ may be interpreted as a probability distribution and sampled with Monte-Carlo methods. \subsection{ Numerical Stabilization} The origin of the numerical instabilities occurring in the calculation of Green functions may be understood by considering free electrons on a two-dimensional square lattice. \begin{equation} H = -t \sum_{<\vec{i},\vec{j}>} c_{\vec{i}}^{\dagger} c_{\vec{j}}. \end{equation} Here, the sum runs over nearest-neighbors. For this Hamiltonian one has: \begin{equation} \langle \Psi_0 | c_{\vec{k}}^{\dagger}(\tau) c_{\vec{k}} | \Psi_0 \rangle = \exp \left( \tau (\epsilon_{\vec{k}} - \mu) \right) \langle \Psi_0 | c_{\vec{k}}^{\dagger} c_{\vec{k}} | \Psi_0 \rangle, \end{equation} where $\epsilon_{\vec{k}} = -2t(\cos(\vec{k} \vec{a}_x) + \cos(\vec{k} \vec{a}_y) ) $, $\vec{a}_x$, $\vec{a}_y$ being the lattice constants. The chemical potential satisfies equation (\ref{chem}) and we will assume $ | \Psi_0 \rangle $ to be non-degenerate. In a numerical calculation the eigenvalues and eigenvectors of the above Hamiltonian will be known up to machine precision, $\epsilon$. In the case $ \epsilon_{\vec{k}} - \mu > 0 $, $ \langle \Psi_0 | c_{\vec{k}}^{\dagger} c_{\vec{k}} | \Psi_0 \rangle \equiv 0$. However, on a finite precision machine the later quantity will take a value of the order of $\epsilon$. When calculating $ \langle \Psi_0 | c_{\vec{k}}^{\dagger}(\tau) c_{\vec{k}} | \Psi_0 \rangle $ this roundoff error will be blown up exponentially and the result for large values of $\tau$ will be unreliable. In order to circumvent this problem, one may do the calculation at finite temperature and then take the limit of vanishingly small temperatures: \begin{equation} \langle \Psi_0 | c_{\vec{k}}^{\dagger}(\tau) c_{\vec{k}} | \Psi_0 \rangle = \lim_{\beta \rightarrow \infty} \frac { \exp \left( \tau (\epsilon_{\vec{k}} - \mu) \right) } { 1 + \exp \left( \beta (\epsilon_{\vec{k}} - \mu) \right) }. \end{equation} Even if the eigenvalues are known only up to machine precision, the right hand side of the above equation for large but finite values of $\beta$ is a numerically stable operation. Although very simple, this example reflects the underlying numerical instabilities occurring in the calculation of the Green functions. We now consider the calculation of \begin{eqnarray} \label{GST} & & G_{\vec{s}} ( \Theta + \tau, \Theta) = B_{\vec{s}} (\Theta + \tau, \Theta) G_{\vec{s}} ( \Theta, \Theta) \; \; {\rm and } \nonumber \\ & & G_{\vec{s}} ( \Theta , \Theta + \tau) = \left( G_{\vec{s}} ( \Theta , \Theta) - I \right) B_{\vec{s}}^{-1} (\Theta + \tau, \Theta) \end{eqnarray} required to compute $G^{>}_{x,y}(\Theta,\tau)$ (see equation (\ref{G>})) and $G^{<}_{x,y}(\Theta,\tau)$ (see equation (\ref{G<})) respectively. The equal-time Green functions, $G_{\vec{s}} ( \Theta, \Theta)$, may be calculated to machine precision \cite{Sandro,Imada,Assaad}. The matrices $B_{\vec{s}} (\Theta + \tau, \Theta) $ contain scales which grow and decrease exponentially with $\tau$. As in the above example, a straightforward multiplication of both matrices will lead to numerical instabilities for large values of $\tau$. Here, the problem is much more severe since the presence of the HS field mixes different scales. In order to circumvent this problem, we propose the following stabilization scheme. Since the trial wave function is a Slater determinant, we can find a single particle Hamiltonian, $H_0 = \sum_{x,y} c^{\dagger}_x (h_0)_{x,y} c_y $, which has $ | \Psi_T \rangle $ as a non-degenerate ground state. The equal time Green functions may then be written as: \begin{eqnarray} \label{GS} \left( G_{\vec{s}} ( \Theta, \Theta) \right)_{x,y} \equiv & & \frac { \langle \Psi_T | U_{\vec{s}}(2\Theta, \Theta) c_x c_y^{\dagger} U_{\vec{s}}(\Theta,0) | \Psi_T \rangle } { \langle \Psi_T | U_{\vec{s}}(2\Theta, 0) | \Psi_T \rangle } = \nonumber \\ \lim_{\beta \rightarrow \infty } \frac { {\rm Tr} \left( e^{-\beta H_0} U_{\vec{s}}(2\Theta, \Theta) c_x c_y^{\dagger} U_{\vec{s}}(\Theta,0) \right) } { {\rm Tr} \left( e^{-\beta H_0} U_{\vec{s}}(2\Theta, 0) \right) } & & = \lim_{\beta \rightarrow \infty } \left( I + B_{\vec{s}} (\Theta,0)e^{-\beta h_0} B_{\vec{s}} (2 \Theta, \Theta) \right)^{-1}_{x,y} \end{eqnarray} The last equality follows after integration of the fermionic degrees of freedom. Inspiring ourselves from the work of Hirsch \cite{Hirsch1} we calculate the time displaced Green functions in equation (\ref{GST}) with: \begin{eqnarray} \lim_{\beta \rightarrow \infty } \left( \begin{array}{cc} I & B_{\vec{s}} (\Theta,0)e^{-\beta h_0} B_{\vec{s}}(2 \Theta, \Theta + \tau ) \\ -B_{\vec{s}}(\Theta + \tau, \Theta) & I \\ \end{array} \right)^{-1} = \nonumber \\ \left( \begin{array}{cc} G_{\vec{s}} ( \Theta, \Theta) & G_{\vec{s}} ( \Theta, \Theta + \tau ) \\ G_{\vec{s}} ( \Theta + \tau , \Theta) & G_{\vec{s}} ( \Theta + \tau, \Theta + \tau ) \\ \end{array} \right) \end{eqnarray} For very large but finite values of $\beta$, we can calculate the left hand side of the above equation by using matrix stabilization techniques developed for finite temperature QMC algorithms. The basic idea behind those numerical stabilization techniques is to keep the different scales occurring in the matrices $B_{\vec{s}}$ separate (for a review see reference \cite{Loh}). This is achieved by decomposing the matrices $B_{\vec{s}}$ into a $UDV$ form where $U$ is an orthogonal matrix, $D$ a diagonal matrix containing the exponentially large and exponentially small scales, and $V$ a triangular matrix. The calculation of the left hand side of the above equation is done in the following way: \begin{eqnarray*} \left( \begin{array}{cc} I & B_{\vec{s}} (\Theta,0)e^{-\beta h_0} B_{\vec{s}}(2 \Theta, \Theta + \tau ) \\ -B_{\vec{s}}(\Theta + \tau, \Theta) & I \\ \end{array} \right)^{-1} = \left( \begin{array}{cc} I & U_1 D_1 V_1 \\ U_2 D_2 V_2 & I \\ \end{array} \right)^{-1} = \end{eqnarray*} \begin{eqnarray*} \left( \begin{array}{cc} V_2 & 0 \\ 0 & V_1 \\ \end{array} \right)^{-1} \left( \begin{array}{cc} \left(V_2 U_1 \right)^{-1} & D_1 \\ D_2 & \left(V_1 U_2 \right)^{-1} \\ \end{array} \right)^{-1} \left( \begin{array}{cc} U_1 & 0 \\ 0 & U_2 \\ \end{array} \right)^{-1} = \end{eqnarray*} \begin{eqnarray} \left( \begin{array}{cc} V_2 & 0\\ 0 & V_1\\ \end{array} \right)^{-1} \left( U_3 D_3 V_3\right)^{-1} \left( \begin{array}{cc} U_1 & 0 \\ 0 & U_2 \\ \end{array} \right)^{-1}. \end{eqnarray} In the above equation, the matrix $D_3$ contains only exponentially large scales since the matrices $\left(V_2 U_1 \right)^{-1} $ and $\left(V_1 U_2 \right)^{-1} $ act as a cutoff to the exponentially small scales in the matrices $D_2$ and $D_1$. Since the other matrices are all well conditioned, the final matrix multiplication is well defined. A convenient choice of $H_0$ is obtained in a basis where the trial wave function may be written as: \begin{equation} | \Psi_T \rangle = \prod _{n = 1}^{N_p} \gamma_n^{\dagger} | 0 \rangle. \end{equation} In this basis, we define $H_0$ through \begin{equation} H_0 \gamma_n^{\dagger} | 0 \rangle = \left\{ \begin{array}{c} - \gamma_n^{\dagger} | 0 \rangle \;\; {\rm if } \;\; \gamma_n^{\dagger} \gamma_n | \Psi_T \rangle = | \Psi_T \rangle \\ + \gamma_n^{\dagger} | 0 \rangle \;\; {\rm if } \;\; \gamma_n^{\dagger} \gamma_n | \Psi_T \rangle = 0 \\ \end{array} \right. \end{equation} (Here, the energy unit is set by the hopping matrix element $t$.) For this choice of $H_0$ values of $\beta t \sim 40$ were well sufficient to satisfy equation (\ref{GS}) within required numerical precision \cite{Note1}. The above numerical stabilization scheme was indeed successful in all examined cases. \section{Evaluation of the charge gap for the two dimensional Hubbard model} We carried out our simulations on $4 \times 4$ to $12 \times 12$ lattices for the two-dimensional half-filled ($\mu = 0$) repulsive Hubbard model (\ref{Hubb}) at $U/t = 4$. Periodic boundary conditions were assumed. A spin singlet ground state of the kinetic energy in the Hubbard Hamiltonian was used as a trial wave function. We test the quality of this trial wave function on a $6 \times 6$ lattice. Figure 1 plots $ \langle \Psi_T |e^{-\Theta H } O e^{-\Theta H } | \Psi_T \rangle / \langle \Psi_T |e^{-2\Theta H } | \Psi_T \rangle $ as a function of the projection parameter $\Theta$ for $ O = S(\pi,\pi)/N = \frac{4}{3N} \sum_{\vec{r}} \exp \left( i \vec{Q} \vec{r} \right) \vec{S}(\vec{r}) \cdot \vec{S}( \vec{0} ) $ (solid circles in Figure 1a) and $ O = E/N = H/N - U/4$ (solid circles in Figure 1b). Here $\vec{Q}= (\pi, \pi)/a$, $\vec{S}(\vec{r})$ is the spin operator on site $\vec{r}$ and $N$ denotes the number of sites. Already for values of $\Theta t = 2.5$, both considered observables have converged within our estimated statistical uncertainty. For comparison, we have plotted $ {\rm Tr } \left( e^{-2 \Theta H } O \right) / {\rm Tr } \left( e^{-2\Theta H } \right) $ for the same observables (triangles in Figure 1). Values of $\Theta t$ at least twice as large are required to obtain approximate ground state results. Another source of systematic errors comes from the discretization of the imaginary time propagation. In Table \ref{table1} the $\Delta \tau$ dependence of the energy and $S(\pi,\pi)$ is given. The data are obtained form the zero-temperature QMC algorithm on a $6 \times 6$ lattice and at $2\Theta t = 5$. The values at $\Delta \tau = 0$ are obtained from a least square fit of the finite $\Delta \tau$ results to the form $a + b (\Delta \tau )^2$. We carried out our simulations at $\Delta \tau t = 0.125$. As may be seen from Table \ref{table1}, this value of $\Delta \tau t$ produces a systematic error contained in the quoted errorbars for the energy and a systematic error of less than $1 \% $ for $ S(\pi,\pi) $. To obtain an estimate of the charge gap, we consider \begin{equation} G(\vec{r} = 0, \tau) = \frac{1}{N} \sum_{x} G_{x,x}(\tau), \; \; \tau > 0. \end{equation} Here, $x$ stands for spin and site indices. Inserting a complete set of eigenstates of the Hamiltonian $H$ in the $N+1$ particle Hilbert space yields: \begin{equation} G(\vec{r} = 0, \tau) = \frac{1}{N} \sum_{n,x} | \langle \Psi_{0}^{N} | c_x | \Psi_{n}^{N +1} \rangle |^2 \exp \left( -\tau \left( E_n^{N+1} - E_0^N \right) \right). \end{equation} where $H | \Psi_{n}^{N +1} \rangle = E_n^{N+1} | \Psi_{n}^{N +1} \rangle $ and $H | \Psi_{0}^{N} \rangle = E_0^N | \Psi_{0}^{N} \rangle $. At large values of $\tau t $, $G(\vec{r} = 0, \tau) \sim \exp \left( - \tau \Delta_c \right) $ where $ \Delta_c \equiv E_0^{N+1} - E_0^N $ corresponds to the charge gap. Our results are plotted in Figure 2. For those simulations we have chosen $\Theta t = 13.5$. Since values of $\tau$ up to $\tau_{max} t = 12 $ were considered, the effective projection parameter is given by: $\Theta_{eff} = \Theta - \tau_{max}/2 = 7.5/t$. As may be seen from Figure 1, this value of the projection parameter is more than sufficient to filter out the ground state from the trial wave function. The solid lines in Figure 2 correspond to least square fits of the tail of $G(\vec{r} = 0, \tau)$ to the form $\exp( - \tau \Delta_c) $ \cite{Note2}. The estimated value of the gap as a function of system size is plotted in Figure 3. A least square fit of the data to the form $a + b/L$, where $L$ denotes the linear length of the lattice, yields $\Delta_c/t = 0.67 \pm 0.02$ in the thermodynamic limit. This value stands in good agreement with the value of the charge gap obtained by Furukawa and Imada \cite{Furukawa}: $\Delta_c/t = 0.58 \pm 0.08$ (solid circle in Figure 3.) As may be seen from the comparison of errorbars, the accuracy of the estimation has been much improved in the present study. \section{Conclusions} We have presented an efficient, numerically stable, QMC algorithm to calculate $T=0$ imaginary time Green functions for Hubbard type models. As a non-trivial test application of this algorithm, we have obtained an accurate estimate of the charge gap for the two-dimensional half-filled repulsive Hubbard model at $U/t = 4$: $\Delta_c/t = 0.67 \pm 0.02$. The algorithm is formulated in the canonical ensemble. Hence, the relation $n(\mu)$ has to be known prior to the simulation. This renders the algorithm hard to use in a metallic state. However, in an insulating state at zero temperature $n(\mu)$ is constant, and generally known, for chemical potentials within the charge gap. In this situation the here presented algorithm proves to be a powerful tool. We illustrate this by considering the two-dimensional Hubbard model. Due to Equation (\ref{factor}), it suffices to know the Green functions at $\mu = 0$ (half-filling) to be able to determine them trivially for any other chemical potential within the charge gap. At $\mu = 0$, we are not confronted with a sign problem due to particle-hole symmetry and the statistical fluctuations do not blow-up exponentially with growing lattice sizes and projection parameters $\Theta$. It is however clear from equation (\ref{factor}) that statistical fluctuations will increase (decrease) exponentially with growing positive values of $\tau$ for $\mu > 0$ ($\mu < 0$). In comparison, finite temperature algorithms in the grand-canonical ensemble, are faced with a sign problem away from $ \mu = 0$. Hence, statistical fluctuations grow exponentially with growing lattice size and inverse temperature. Away from $\mu = 0$, it is thus extremely hard to extrapolate any zero temperature result from the finite temperature grand-canonical algorithms for large lattice sizes. This renders the here presented algorithm a powerful tool for the study of the metal-insulator transition from the insulator side \cite{Assaad1}. \section*{Acknowledgements} F.F. Assaad thanks the JSPS for financial support. The numerical calculations were carried out on the Fujitsu VPP500 of the Supercomputer Center of the Institute for Solid State Physics, Univ. of Tokyo. This work is supported by a Grant-in-Aid for Scientific Research on the Priority Area "Anomalous Metallic State near the Mott Transition" from the Ministry of Education, Science and Culture, Japan.
\section{Introduction} Our vantage point inside the Milky Way disk offers us the possibility of a unique insight into the structure of at least our galaxy; however this same location causes many unique problems. It is undoubtedly useful to compare the Milky Way with other, more distant spiral galaxies, but this rarely happens on an equal footing: sometimes the Milky Way serves as a local well-understood calibrator for the external galaxies, other times it is the other galaxies which provide inspiration and guidance necessary for us to be able to interpret the observations of our own Galaxy. Though the study of the dynamics of the inner regions of the Milky Way may be said to be still in the `borrowing from other galaxies' phase, the many data being gathered make it likely that eventually we will be able to `give' as well. This review covers the mounting evidence that the center of the Galaxy harbours a bar with a size of a few kpc. Early evidence for a bar (\S2) was championed mostly by de Vaucouleurs, but received little following until interest was rekindled about five years ago by results from near-infrared surveys. Since then, many detections of a barred distortion, broadly consistent with each other (at least as regards direction of the bar major axis) have appeared. They fall into two broad categories, those based on photometric data (surface photometry and star counts, \S3) and those based on kinematics of stars and gas (\S4). More recently, the gravitational microlensing searches in the direction of the Galaxy bulge have turned up many more events than had been originally expected based on a simple axisymmetric model for the Milky Way. A bar may significantly enhance these expected rates, and may well be required to explain the microlensing data (\S5). This review ends with a wish list of some observations which might help constrain the bar parameters in the future (\S6). \begin{table} \caption{Properties of the Galaxy with a bearing on its Hubble type, according to de Vaucouleurs (1970). Each property/morphological type pair at stage Sbc is assigned a score of $+1$ (good agreement), 0 (indifferent), or $-1$ (conflict).} \label{tabdv} \begin{centering} \begin{tabular}{lccccccc} Criterion & (a) & (b) & (c) & (d) & (e) & (f) & \multicolumn{1}{c}{Sum} \\ \tableline A(s) & $-$ & $-$ & $-$ & $-$ & & $-$ & $-5$ \\ AB(s) & $-$ & $-$ & $-$ & $-$ & + & $-$ & $-4$ \\ B(s) & $-$ & $-$ & $-$ & & & $-$ & $-4$ \\ B(rs) & & + & + & & & + & +3 \\ B(r) & & + & & + & & & +2 \\ AB(r) & + & + & & & & & +2 \\ A(r) & & & & $-$ & & $-$ & $-2$ \\ A(rs) & & & + & $-$ & + & $-$ & 0 \\ AB(rs) & + & + & + & & & + & +4 \\ \tableline\tableline \end{tabular} \end{centering} \tablenotetext{a}{High spiral arm multiplicity} \tablenotetext{b}{Inner ring diameter of 6kpc} \tablenotetext{c}{Broken ring structure} \tablenotetext{d}{Radio structure of the nucleus} \tablenotetext{e}{Yerkes spectral type} \tablenotetext{f}{Non-circular {\sc Hi} motions} \end{table} \section{Early Evidence} De Vaucouleurs (1964, 1970) early on suggested that the Milky Way was in fact a barred spiral. His argument relied on comparing many morphological features of the Milky Way with other spirals of different revised Hubble types, the crucial step in this analysis being to associate the {\sc Hi} feature known as the 3-kpc arm (also interpreted in terms of a bar by Kerr 1967) as part of a broken {\sc Hi} ring. For a list of properties, he gave each of the subtypes (r,rs,s) and (A,AB,B) a score reflecting their goodness of fit for the Milky Way, and obtained an overall score by a simple unweighted sum (see Table~1). The best fit was the completely mixed type SAB(rs)bc: a galaxy with fairly weak rings and a bar with not-quite grand design spiral structure. It is interesting, and reassuring, that this conclusion does not appear to be dominated by a single column in Table~\ref{tabdv} (though three of the six diagnostics pertain to the 3-kpc expanding arm). There thus appeared to be a good case for taking this suggestion seriously. However, much of the effort in understanding Galactic structure in subsequent years was focused on the problem of maintaining spiral density waves, and the idea that the Milky Way had a bar fell out of fashion. Only in the last five years has it returned into favour. \section{Photometric Evidence} Because the Galaxy is virtually edge-on, we cannot see a bar directly. However, unless the bar happens to be aligned along or perpendicular to the sun-Galactic center line, a bar will create systematic differences between points at equal and opposite longitudes: if the major axis of the bar lies in the first quadrant, for instance, (longitude $0<\ell<90^\circ$) then objects in that quadrant will on average be closer to the sun than those in the fourth quadrant. Such effects can show up both in surface brightness and in star counts. \subsection{Surface photometry} Blitz \& Spergel (1991), in their search for non-axisymmetric structure in the Milky Way, analysed the near-IR data of Matsumoto et al.\ (1982) and showed that there were indeed systematic differences between positive and negative longitudes near the Galactic center. They showed that these could be understood as a perspective view of a bar: the near side (in the first quadrant) would appear more vertically extended on the sky than the far side, and the surface brightness of the near side should be greater, both as observed. Furthermore, in the innermost regions the asymmetry in surface brightness is actually reversed, and this feature too is reproduced in the data. \begin{figure} \plotone{dwekbar2.ps} \caption{The bar models `G2' (solid lines) and 'E3' (dashed) of Dwek et al. (1995) projected onto the Galactic plane. Contours are spaced at factors of 3 in mid-plane density; the outermost contour corresponds to $3\times10^6L_\odot/\rm kpc^{-3}$. The position of the sun ($R_0=8$kpc) is indicated, and galactic longitude increases counterclockwise from the right.} \end{figure} Superior data from the DIRBE experiment on the COBE satellite have confirmed and sharpened this result. After dereddening the near-IR data, Dwek et al. (1995) derive a best-fit model `G2' for the emissivity of the bar of the form \begin{equation} \rho(x,y,z)\propto e^{-s^2/2} \end{equation} where (assuming the sun is 8kpc from the Galactic center) \begin{equation} s^4=\left[\left(x\over1490\rm pc\right)^2+\left(y\over580\rm pc\right)^2\right] +\left(z\over400\rm pc\right)^4 \end{equation} and $(x,y,z)$ are Galactic coordinates rotated by $13.4^\circ$ about the Galactic minor ($z$-) axis. This functional form implies an ellipsoidal shape for the bar projected onto the galactic plane, but makes boxy bulge isophotes as seen from earth, as observed. (To some at this conference, the boxy isophotes are already strong evidence for a bar!). The axis ratios are 2.6:1:0.7. There are still uncertainties in this deprojection, which was derived as a least-square fit to a fairly restricted set of models---it will be hard to do better given the usual problems associated with recovering a three-dimensional emissivity from a two-dimensional surface brightness map. For instance, model `E3', a triaxial version of Kent's (1992) modified Bessel function model fits almost as well as G2, but with major axis position angle $40^\circ$ (this model does have the unsatisfactory feature that the z:y axis ratio is greater than 1). Further uncertainty arises from the correction for the disk contribution to the surface brightness. A sketch of both bar models is shown in Figure~1. \subsection{Star counts} Counts of individual objects also reveal left-right asymmetries of the Galactic center region. All such data sets show an effect in the same direction: objects at positive longitudes appear brighter and therefore are presumably closer. These data sets include SiO masers (Nakada et al. 1991, Izumiura et al.\ 1994), IRAS AGB stars (Weinberg 1992), IRAS Miras (Whitelock \& Catchpole 1992), the OGLE red clump stars (Stanek et al. 1994), and OH/IR stars (Sevenster, this volume). Furthermore, the globular clusters also appear to show a bar-like distortion (Blitz 1993). Typical magnitude offsets are 0.2--0.5, the best-measured one being that of the OGLE group ($0.37\pm0.03$). For comparison, the Dwek et al. (1995) G2 model would allow at most a 0.2 magnitude offset between brightness of objects at positive and negative longtudes within $6^\circ$ of the galactic center (Figure~2). \begin{figure} \plotone{dwekdm.ps} \caption{The average magnitude offset between objects at galactic coordinates $(\ell,b)$ and $(-\ell,b)$ in the Dwek et al. bar model. Different curves correspond to different latitudes: (top to bottom) $b=0,2,4,6,8,10^\circ$.} \end{figure} While all these studies agree on the sign of the asymmetry, the agreement mostly ends there: the magnitude offset between positive and negative $\ell$ varies considerably from survey to survey (though part of the effect may be due to small number statistics in some of these data sets, and different depths of the different samples). Also, the results of the bulge surface photometry imply a very much smaller bar than the IRAS AGB counts of Weinberg: the former extends out to longitudes of about $10^\circ$, and the latter out to about $40^\circ$. It is therefore quite possible that the Milky Way in fact contains {\em several} bars. It is interesting to note that, had we not known about bars in other galaxies, we might not have chosen to interpret these left-right asymmetries in such terms. On the sky, the asymmetries suggest a lop-sided ($m=1$) distortion instead, and it might have seemed far-fetched to attribute these to perspective effects of an inherently $m=2$ bar. \section{Kinematic Evidence} Bars also show up as kinematic distortions of the velocity fields of stars and gas, since the closed orbits in a pattern-rotating barred potential are no longer circular, but rather elongated along or perpendicular to the bar. Resonances (chiefly inner and outer Lindblad and corotation) affect the orbit structure profoundly. Some closed orbits are self- or mutually intersecting, making them unsuitable as gas orbits and consequently generating gaps and shocks in the gas distribution (see the review by Athanassoula in this volume), and the distribution of stellar orbits follows similar behaviour. Unlike photometric signatures (except perspective effects discussed above), these kinematic effects of a bar are also visible in edge-on systems, and so provide a means of detecting bars in such galaxies. Both the kinematics of gas and stars could reveal evidence for a bar in the Milky Way, but each have their problems when it comes to quantifying bar parameters. Gas, because it tends to dissipate down to the closed, non-intersecting orbits, delineates the orbit structure and hence the potential most clearly, but the most striking features are associated with the resonances where hydro-dynamic effects are a major factor. It is still very difficult to model all the relevant processes at these locations well. Stellar orbits, on the other hand, are dominated by gravitational forces, but because of the absence of dissipation the accessible orbits are much more varied. It is still an unsolved problem to derive the gravitational potential from observed radial velocity distributions in stationary elliptical galaxies, and the barred galaxy problem, which also involves unknown figure rotation, is even more complicated. Therefore, though it is possible to rule out axisymmetric models on the basis of kinematic data, producing a unique bar model is a considerably harder problem. \subsection{Gas kinematics} The distribution of gas within $5^\circ$ of the Galactic center is complicated. Significant features for our purposes are: \begin{itemize} \item Large forbidden CO and {\sc Hi} velocities \item A fast outwards decline in {\sc Hi} tangent point velocities \item A {\em very} lopsided CO distribution \item A dramatic change in the CO kinematics near longitudes $+1.7$ and $-1^\circ$. \item A tilted (by $\sim7^\circ$ projected onto the sky) {\sc Hi} (and maybe CO) plane. \end{itemize} The CO velocity structure, from the Bell Labs survey (Bally et al.\ 1987, 1988) is shown in Figure~3. \begin{figure} \plotone{parall.ps} \caption{Left: the distribution of CO emission in longitude and radial velocity, from the Bell Labs.\ survey. Emission at $b=-3'$ is shown. Contours are drawn at brightness temperatures of 1,2,4,8 and 16 K. Note the parallelogram-shaped envelope of the emission. Right: a model for the parallelogram, from Binney et al.\ (1992). The cusped orbit is viewed from the direction of the thick arrow, and gas streams around the orbit as indicated.} \end{figure} The forbidden velocities (negative velocities in the first quadrant, positive ones in the fourth) imply non-axisymmetry, assuming the gas to be a dynamically cold tracer of the potential. However, per se they say little about the nature of this deviation from circularity: in particular, local expanding features in the gas may cause features similar to those observed (e.g. Oort 1977, Uchida et al. 1994). So far, no coherent {\em dynamical} model has been formulated which addresses all observed features (but see Weinberg's paper in this volume). However, several analyses have focused on subsets of these observations. All these investigations have centered on single-bar models, though reality may well be more complex. Liszt \& Burton (1980) have modelled the kinematics of the {\sc Hi} in the central few kpc as a tilted, elliptically streaming disk (an earlier fit as an expanding disk was equally succesful, if less plausible). Their model succesfully fits the observed distribution of {\sc Hi} in position on the sky and radial velocity, though it offers no dynamical origin for this disk. If the ellipticity is caused by a bar, then the least satisfactory aspect of this model is the absence of pattern rotation. It seems plausible, though, that a pattern speed could fairly simply be accomodated in such a kinematic model. Mulder \& Liem (1986) attempted to construct a global model for the {\sc Hi}. Using non-selfgravitating hydrodynamical simulations (pioneered in this context by Sanders \& Huntley 1976), they showed that a multitude of kinematic features in the Galactic {\sc Hi} could be explained with a simple model in which a gas disk is forced into a quasi-steady flow by a simple model for a weak, rotating bar. In particular, the 3kpc arm could be identified with shocked material near the inner Linblad resonance, while the sun's position near corotation (implying quite a slow pattern speed) explained the three nearby spiral arms. Forbidden velocities in the central few degrees could also be accounted for. Their striking results, however, appear to have received relatively little interest at the time. Binney et al.\ (1992) concentrated on the distribution of the CO and other molecular gas at $b=0$ (Fig.~3), and interpreted it in terms of a dynamical model in which the gas is allowed to move on closed, non-intersecting orbits only. No attempt was made to address the tilt of the inner plane. They identify the striking parallellogram shape of the orbit with the smallest orbit outside the inner Lindblad resonance which does not intersect itself---gas further in will strongly dissipate kinetic energy and end up in inside the ILR on an `$x_2$ disk'. The parallelogram orbit is cusped, and seen from a fairly narrow range of angles, its projection into the longitude-velocity plane takes on the observed shape. Because this orbit is strongly affected by the resonances, the pattern speed of the Binney et al.\ model is very well constrained, with corotation around $2.4\pm0.5\,$kpc. Furthermore, the parallellogram projection of the orbit only appears from viewing angles of the bar about $16^\circ$ off end-on. The distribution of {\sc Hi} at larger radii is consistent with the closed orbits outside corotation, as is the radial dependence of the model bar density with that of the observed bulge light. In spite of these successes, a worrying aspect of this model is the left-right asymmetry of the parallelogram. The data show such an effect in the sense expected from perspective, but it is much more pronounced than expected from the model. Rigorous modelling of the observed asymetry (the cusps of the orbit appear at longitudes $\sim+1.7^\circ$ and $-1^\circ$) implies that the cusps of the parallelogram orbit lie at radius $\sim R_0/5\simeq1.6$kpc. An orbit of this size would have to be aligned within $6^\circ$ of the line of sight, and would have to be very slender if we were indeed viewing it down its sides. A more plausible, if perhaps less elegant, explanation invokes some lopsidedness to the central kinematics, which spoils the perspective of the bar orbit. Such a component may be needed anyway to explain the rather large velocity difference of the gas deduced to lie near the cusps: this gas should have zero velocity in the bar frame. The dynamics of the tilt of the inner gas are a puzzle. It may have consequences for the bar analysis: when Liszt \& Burton (1980) restrict their model to the Galactic plane (rather than the tilted inner disk plane), velocity crowding of the gas mimicks the observed distribution of CO very nicely. It therefore appears that the identification of the parallellogram with a specific CO orbit is not unique, and a more detailed consideration of the CO distribution out of the galactic plane is required. (Initial suggestions by Blitz \& Spergel (1991) that the stellar emission is tilted in the same direction as the gas were shown by the COBE data to have been an artefact of extinction. The stellar distribution is consistent with being aligned with the Galactic plane). Weiner \& Sellwood (1995) have concentrated on fitting the kinematics of the {\sc Hi}, particularly the sharp falloff of the tangent point velocity, outside longitudes $4^\circ$. They use a hydrodynamic code to model the steady-state behaviour of the gas. Their results appear inconsistent with the findings of Binney et al.: they find that only bars seen over $30^\circ$ off end-on can generate forbidden velocities over a sufficiently large longitude range. Their best-fit model also has a significantly larger corotation radius of 3.6kpc. The differences between the various analyses of the gas kinematics partly reflect differences between the kinematics of the different tracers, possibly due to nongravitational effects, but to some extent also is an indication of the collisionality of the interstellar medium: it remains a gross simplification to model gas as perfect tracers of the closed non-intersecting orbits in a smooth, pattern-rotating potential. For instance, the lop-sided distribution of the central gas distribution is possibly a transient feature (e.g., a fluctuation associated with the relatively small number of large clumps in the central few 100pc, or the result of a dynamical instability or interaction) whose amplitude raises concerns about fitting equilibrium bisymmetric models to the dynamics. An investigation by Jenkins \& Binney (1994) shows that stochastic processes in the gas distribution will indeed cause lopsidedness, but they have difficulty reproducing effects as dramatic as those observed. Quite possibly, self-gravity or low-temperature cooling (neither of which is included in their calculations) can make a substantial difference here. Future refinements of the analyses of the observed gas dynamics may well be inspired on observations of the CO distributions in other barred galaxies (see reviews by Turner and Kenney in this volume), which may establish when molecular gas does and does not trace the closed orbits allowed by the potential. \subsection{Stellar kinematics} Given the possible problems with the dynamics of the gas, in how far can stellar dynamics help? At the moment, the answer is, unfortunately, not very much. The large velocity dispersions in the bulge region wash out signatures of non-axisymmetry, which only large numbers of stars sampled at a range of longitude or integrated-light velocity distributions (see Kuijken \& Merrifield 1995 or Merrifield, this volume) can overcome. Apart from the difficulty of getting sufficiently detailed observations, there is also a theoretical bottleneck: we do not know what the velocity distributions in realistic galactic bars actually look like, because there are large families of possible combinations of stellar orbits which can be combined to make the same bar. Whereas gas modelling can be simplified by considering closed orbits, this constraint is not available in the case of stars. Ideally, it should be replaced by a further observational phase-space measurement: distance down the line of sight and/or proper motions. In any case, just about the simplest axisymmetric model that can be constructed for the bulge, an oblate isotropic rotator, appears to fit all available stellar-kinematic data (Kent 1992), including recently published M giant samples (Blum et al. 1995). This good fit is not evidence against a bar, but rather an illustration of the difficulty of detecting a bar in the stellar kinematics of the bulge. The strongest feature in radial velocity data that argues in favour of a bar is the bimodality of the OH/IR stars: in addition to a hot `bulge' population, the central degree or so contains quite a cold stellar disk, whose kinematics are similar to those of the inner CO gas (Lindqvist et al. 1992). It is possible that these stars were formed from the gas that lives inside the inner Lindblad resonance (an `$x_2$ disk'). The most detailed model constructed for the Milky Way's stellar bar is that of Zhao (this volume). Analysis of the Spaenhauer et al.\ (1991) sample of stars with proper motion in Baade's Window (Zhao, Spergel \& Rich 1994) shows possible signatures of triaxiality (vertex deviations of metal-weak and metal-poor stars incompatible with axisymmetry), but since the sample is small the statistical weight of this study is rather low. Similar analyses of larger samples in different parts of the bulge currently offer the best hope of understanding the bar dynamics from stellar kinematics. Long-range perturbations of the stellar kinematics by the quadrupole field of a bar may also be detectable. Perturbation formulae for the stellar velocity field, as well as the velocity dispersions, in a barred, pattern-rotating potential, have been derived by Kuijken \& Tremaine (1991). Weinberg (1994) has shown that the resonances of a bar with a decreasing pattern speed will trace out a characteristic signature across a disk, and he finds some evidence in the kinematics of old K giants for such a feature. \subsection{The pattern speed from stellar kinematics} A particularly important product of stellar kinematics might be the measurement of the pattern speed $\Omega_p$ of the bar. Such measurements can be made in model-dependent ways by identifying certain morphological features (typically of the gas) with resonances, or less so via an integral constraint derived from the continuity equation (Tremaine \& Weinberg 1984, TW). The TW method involves integration along a given line of the velocity component perpendicular to it. It was originally formulated for application to moderately inclined external barred galaxies, in which case it requires measurement of mean radial velocity along a line parallel to the major axis. In that form it is inapplicable to edge-on galaxies such as our own, but two modifications are: the first involves integration of heliocentric radial velocities around the galactic equator (Kuijken \& Tremaine 1991) and the second integration of transverse velocities down a single line of sight near the Galactic center. Neither variant is currently practical, however: the second requires accurate distances and proper motions at $b=0$, whereas the first would rely on full longitude coverage in the densest parts of the galactic plane, with complete radial velocity coverage. Nevertheless, future large near-infrared surveys may one day allow these measurements to be made. \section{Gravitational Microlensing Evidence} Microlensing of stars by foreground objects which pass at very small projected impact parameters has recently developed from an elegant curiosity to a new tool in galactic structure research. As shown by Refsdal (1964), if a foreground object of mass $m$ at distance $x$ from us passes within a radius \begin{equation} R_E(x)=2\sqrt{Gmx(1-x/L)}/c\propto m^{1/2} \end{equation} from the line of sight to a source at distance $L>x$, the source will be magnified (`microlensed') by a factor $>1.34$. Since in general the lens will move with respect to the line of sight, the brightening will only last for a certain time, typically of the order of 1-100 days (depending on the lens mass). The average number of lenses in the `microlensing tube' $R(x)<R_E(x)$ is called the optical depth $\tau$, and depends on the number density $\nu$ of lenses along the line of sight: \begin{equation} \tau=\int_0^L\nu(x)\pi R_E(x)^2dx\propto m^0\overline\rho \end{equation} where $\overline\rho$ is a mean mass density in lenses. $\tau$ therefore depends only on the mass density in lenses, not on the masses of individual lenses (but the detectability does depend on $m$ via the timescale of typical events). Whereas the detection rate for microlensing towards the Magellanic clouds may be disappointingly low (Alcock et al.\ 1995a), the `control experiments' towards the Galactic bulge have surprisingly turned up many more events than had initially been expected: the optical depth to the average bulge star is about $3\times10^{-6}$ (Udalski et al.\ 1994, Alcock et al.\ 1995b). Along the line of sight towards Baade's Window, a double exponential disk can produce at most $\tau_D<1.2\times10^{-6}$, with a more likely number being less than half that (Fig.~4). \begin{figure} \plotone{taudisk.ps} \caption{The optical depth of bulge stars to microlensing by a double exponential disk $\rho\propto \exp(-R/a-z/H_d)$, constrained to produce a maximum rotation speed below 180km/s. The sun symbol shows the optical depth due to a less than maximal disk, consistent with the measurements of Kuijken \& Gilmore (1991).} \end{figure} The early calculations were based on axysymmetric models, and on the assumption that the bulge stars were only lensed by foreground disk and halo objects. It was later realized (Kiraga \& Paczinski 1994) that bulge stars are so common that lensing of a far-side bulge star by one on the near side contributes a significant signal ($\tau_B\sim0.7\times10^{-6}$ if one uses the Kent model for the bulge). This signal is enhanced further if the bulge is extended along our line of sight, for then the near-side stars are in a fatter part of the microlensing tubes for lensing of the far-side bulge stars. The effect can be as much as a factor of two if the bulge has the axis ratio of the Dwek et al.\ (1995) model, raising the optical depth to bulge sources to around $1.2\times10^{-6}$. The numbers are still a little low, and larger numbers of microlensing events will have to be analysed before it is clear whether there still is a problem or not. Further constraints on bulge-bulge lensing can be derived by searching for a systematic offset between the (unmagnified) brightnesses of lensed stars with the general population. If far-side bulge sources are systematically lensed more often than near-side ones, the lensed sources should be systematically fainter. The magnitude of the offset can be used to constrain the axis ratio and orientation of the bar (Stanek 1995, and this volume). \section{Conclusions and Wishlist of Further Observations} It is clear that a variety of lines of evidence point towards the existence of non-axisymmetric structure in the central few kpc of the Milky Way. Equally impressive is the lack of evidence to the contrary! While the precise details have not yet been characterised, rapid progress is being made, partly driven by the need to understand the new microlensing data. Major stumbling blocks at the moment are the difficulty of realistically simulating hydro\-dynamical processes. In conclusion, it seems useful to compile a list of observations which may help pin down the nature and parameters of the bar. These include: \begin{itemize} \item {\em To see the bar in stellar kinematics.} Proper motions of samples of stars throughout the bulge will greatly help define the orbit structure, and hence the gravitational potential and pattern speed of the bar. \item{\em An optical depth map of the bulge region.} As shown by Kiraga (1994), such a map provides an entirely separate constraint on the mass distribution in the central regions. \item{\em evolutionary history of the bulge as traced by stellar abundances and their ratios.} Such data can be used to constrain the star formation history of the bulge/bar, and combination with kinematic data ultimately will allow the evolutionary relation between the bar, bulge (if indeed they are separate) and disk to be addressed. \item{\em Self-gravitating hydrodynamic simulations of gas flow in barred potentials} will help address issues related to central lopsidedness, stability and possible tilts. \end{itemize}
\section{Introduction} This paper is part of a series of papers describing the construction and maintenance of a quasi--inertial reference frame in both the radio and optical domain. Extensive work on the radio reference frame has been accomplished in the last 5 years and published in a series of papers I--VII (see Johnston et al. \markcite{rmRORF} 1995 for further details). A rigorous global new reduction of all then applicable Mark III VLBI radio observations has been used to construct a radio reference frame of 560 sources from first principles (Johnston et al. \markcite{rmRORF} 1995). Based on these results a list of defining and candidate sources has been provided to the IAU Working Group on Reference Frames (IAU, \markcite{rmIAUWG} 1995). While a dense radio frame with an accuracy level of 1 milliarcsecond (mas) for most of the source positions is now in place, optical observations on a 30--50 mas level are available for only a fraction of these sources. Previous results already have shown the deficiencies of the currently used optical FK5/J2000 reference frame, with deviations from a uniform inertial reference system as large as $\approx 200$ mas at the current epoch. The Hipparcos astrometry satellite mission will soon provide a new optical system on the 2 mas level at the Hipparcos mean epoch ($\approx 1991.5$), and 2 mas/year in proper motion error, but this instrumental system must be linked to the radio system to become quasi--inertial. Most of the primary optical reference objects (stars) are bright, both in the FK5 (3 to 6 mag) and the Hipparcos (5 to 9 mag) catalogs, while the optical counterparts of the extragalactic sources are optically faint (the majority in the range 17 to 21 mag). Due to the relatively low quantum efficiency of photographic astrometry, which requires long exposure times on large telescopes in good seeing, progress in the optical observations has been slow. The use of CCD detectors has dramatically improved this situation, because they allow smaller telescopes which have greater availability and more objects can be observed due to much shorter exposure times. Not until recently has the fieldsize of CCD's became large enough to contain a sufficient number of reference stars for precise astrometry. This paper outlines the reduction procedure in detail and gives results for a representative subset of the observed sources. It is a pilot investigation to assess the astrometric properties of these telescopes and the capabilities of this technique. The first successful attempt to use CCD's for this project has been made earlier (de Vegt \markcite{rm1CCD} et al. 1987), although it was severely limited by the lack of reference stars in a tiny 2' by 3' field of view. In Section 2 we discuss the telescopes, CCD's and observations and in Section 3 the reference star data. Section 4 describes the reduction procedure, while results are presented in Section 5. An accuracy estimate of the procedure and comparison with other investigations is made in Section 6. \section{OBSERVATIONS} \subsection{The Telescopes} The 0.9 m (36 in) Kitt Peak National Observatory (KPNO) telescope is a Ritchey--Chr\'etien system with an additional 2--element field corrector about $250 \ mm$ before the focal plane. This gives a large ($\approx 1\deg$) flat field of view, which also gives the offaxis guide scope good image quality. Frequent focus measurements and the use of the correlation between focus setting and telescope temperature ensured optimal image quality for all object frames. The image quality at the KPNO instrument was found to be uniformly good over the entire field of the CCD. The 0.9 m (36 in) Cerro Tololo Inter--American Observatory (CTIO) telescope is a cassegrain system without a field corrector. At the edge of the CCD frame optical aberrations are visible and it is difficult to achieve a well focused CCD frame with round images over most of the chip area due to instabilities in the mirror--supporting structure. We also had occasional guiding problems. On the positive side the longer focal length of the CTIO 0.9 m with its better sampling, makes it more suitable for structure analysis of the objects than the KPNO 0.9 m. Getting enough reference stars for the astrometric link was found to be the bottleneck with the CTIO 0.9 m. Properties of both telescopes and the CCD's used are summarized in Table 1. \subsection{The Data Acquisition System} Both telescopes use the same type of Tektronix 2K CCD chip, with square pixels of size $24 \ \mu m$ and a filling factor of 100\%, with a different camera controller. For good astrometric results, a large S/N ratio is required, whereas an optimized digitization at the background level is of minor importance. In order to cover the large magnitude range between the reference stars and the extragalactic sources, a large gain of 8.2 was chosen for the KPNO instrument to utilize the full dynamic range, including the full well capacity of the chip. For the CTIO camera, a gain of 3.3 was sufficient because of the larger digitization range (16 bit) available with that camera controller. The readout time for the KPNO instrument was well over 2 minutes. The new ARCON controller at CTIO allowed a faster readout of the full frame in about 70 seconds with two readout amplifiers. All frames have been taken in a red spectral bandpass. A Gunn r filter was used for the long exposure frames to record the extragalactic objects. For each object, additional short exposure frames have been taken in order to get unsaturated images of the reference stars (12 to 14 mag). With the KPNO 0.9 m, most of these short exposure frames have been taken with a narrow (FWHM = 12 nm) filter centered near $H\alpha$. The better sampling of the CTIO instrument allowed the use of the same Gunn r filter for the short exposure frames because the flux of the secondary reference stars was spread out over more pixels, thus avoiding saturation. The IRAF software system was used for the data acquisition at both sites. \subsection{Observation Procedure} A summary of the observing runs is presented in Table 2. Between April and October 1994 the dome seeing was improved at the 0.9 m KPNO telescope. In the first observing run at each site many calibration and test frames were obtained in order to evaluate the astrometric quality and possible systematic errors of the instrumentation, as well as to determine the best observing strategy. The second runs at each telescope were pure production runs, taking at least 2 long and 2 short exposure frames per field. Depending on the brightness of the sources, an exposure time of 200 to 900 seconds was used for the deep frames and 40 to 120 seconds for the others. At the KPNO telescope a typical deep frame covers an astrometrically useable range of $14 \ to \ 20^{m}$, while the short exposures cover $10.5 \ to \ 16.5^{m}$. At CTIO the corresponding ranges are $13 \ to \ 20^{m}$ and $11 \ to \ 18^{m}$. All frames were taken within 1 hour of the meridian and at least $30\deg$ away from the Moon. In fields of high galactic latitude the density of the secondary reference stars are usually not sufficient for a good astrometric reduction. In those cases a mosaic of short exposures (2 x 2 frames), centered on the object and shifted by 500" in $x$ and $y$ were obtained at KPNO. The overlap of 41\% in area allows a rigid tie to the central frames and the area covered for potential reference stars was increased by a factor of 2.5. Because of the much smaller useable field size of the CTIO 0.9 m, a similar 2 x 2 mosaic with shifts of 240" in x and y were obtained for all fields at that telescope. This is an overlap of 50\% by area. A few additional calibration fields were taken with at least a 3 x 3 set of frames of 2--3 minutes exposure time. On some nights, additional sets of short exposure (10, 20, 40 seconds) frames were obtained in order to investigate the limits set by atmospheric turbulence on astrometric accuracy. Results will be published elsewhere (Zacharias \markcite{rmZatm} 1996). It was necessary to have two observers present, in order to obtain online quality control. On each deep frame, the object was identified and radial profile and contour maps were generated. A few objects were found to be optical doubles on the 2 to 7 arcsecond level, most likely due to foreground stars. Depending on the seeing, these objects were skipped or the exposure time was adjusted, if required. All frames were checked for focus and overall image quality. The short exposures in addition were checked for saturation of the reference star images and the exposure time was adjusted accordingly. \section{Reference Stars} The currently used primary reference system is the IRS (International Reference Stars), (Corbin \& Urban \markcite{rmIRSp} 1990, Corbin \& Warren \markcite{rmIRSc} 1991). The IRS gives positions and proper motions in the FK5/J2000 system for approximately 36,000 stars in the magnitude range of $V \approx 6 \ to \ 9$, nearly uniformly distributed on the sky. Nearly all radio source fields from our candidate list (Johnston et al. \markcite{rmRORF} 1995) already were observed with modern high precision astrographs. The northern hemisphere plates were taken with the 23 cm Hamburg Zone Astrograph (ZA) (de Vegt \markcite{rmCDS} 1978) and the southern hemisphere plates were obtained with the yellow lens (BY) of the U.S. Naval Observatory 8 in Twin Astrograph (Routly \markcite{rmTA} 1983) from the Black Birch Astrometric Observatory (BBAO) in New Zealand. The field sizes of these instruments are $ 6\deg \times 6\deg$ (ZA) and $ 5\deg \times 5\deg$ (BY) respectively. Both instruments use $6^{m}$ objective gratings in order to obtain diffraction images of bright stars for high precision position measurements. The range of magnitudes covered is $V \approx 5 \ to \ 14$ in 15--minute exposures on microflat 103aG emulsion. Both instruments have 2 meter focal length, corresponding to a plate scale of about $100"/mm$. All IRS and Hipparcos Input Catalog (HIC) stars (Turon et al. \markcite{rmINCA} 1992) have been measured on these plates together with all faint stars to the plate limit in an area of $1\deg \times 1\deg$ centered on the radio source position. These stars in the magnitude range of $V \approx 11 \ to \ 14$ serve as secondary reference stars for the reduction of the CCD frames. In addition, the 0.5 m Lick Carnegie Astrograph (LA) was used to provide a denser net of secondary reference stars in selected fields. The LA has a field size of $ \approx \ 3\deg \times 3\deg$ on $240 \ mm \ \times \ 240 \ mm$ plates but a limiting magnitude of $V \approx 15.5$ in 30--minute exposures with a plate scale of $55"/mm$. For many fields on the southern hemisphere, plates from the ESO Schmidt telescope are available, which often show measurable images of the radio sources. All Schmidt plates provide at least a set of tertiary reference stars in the magnitude range of $R \approx 14 \ to \ 18$. Measuring of all those plates is in progress at Hamburg Observatory with the HAM--I machine (Winter et al. \markcite{rmW1} 1992, Winter \markcite{rmW2} 1994, Zacharias et al. \markcite{rmZ161} 1994). An accuracy of $\approx 0.8 \ \mu m$ per coordinate is obtained for the measurement of a single image on a good astrograph plate. For the present pilot study only a subset of CCD observations were used to obtain positions of the optical counterparts of the extragalactic sources. A publication for all fields is in progress and will be based on the Hipparcos catalog. Ultimately all CCD frames of the 0.9 m telescopes depend entirely on the astrograph observations, which provide the high precision secondary reference star positions. The Guide Star Catalog (GSC) is not precise enough for this project and results from CCD transit circle instruments are premature at the moment. The future Tycho catalog alone will not be dense enough to provide a good reduction of the current CCD observations. \section{Reduction Procedure} \subsection{Calibration of Raw CCD Frames} The standard IRAF software package (version 2.10, NOAO, Univ. of Arizona, Tucson) was used for the initial reduction steps of the raw CCD data. About 20 bias frames were combined for a master Zero for each night. Most of the KPNO frames were calibrated with flatfields derived from a large number of object frames from a group of nights in which the dust grain pattern on the filter and dewar window remained constant. CCD frames taken with the narrow filter at KPNO and all CTIO frames were calibrated with twilight flats. \subsection{Pre--processing Statistics} Statistical information such as noise characteristics of the frame and full width half maximum (FWHM) values of the image profiles were obtained from all calibrated images using standard IRAF commands. An approximate pixel position and count rate of the extragalactic objects were obtained from radial profile fits, as well as relative position offsets for the short exposure frames. This information is read by our reduction programs for an automatic data handling of multiple frames and fields. \subsection{Determination of $x,y$ Coordinates} In order to obtain centroid positions of star images ($x,y$ coordinates) two different software packages were investigated for comparison. First, the IRAF/DAOPHOT routines provided three different sets of $x,y$ coordinates: \ a) a center of mass position, \ b) a 1--D gaussian fit to the $x,y$ marginal distribution of the pixel data, and \ c) a point spread function (PSF) fit after subtraction of a gaussian. Only options a) and b), provided by the routine {\em phot}, include an error estimate on the derived $x,y$. Option c) is obtained with {\em allstar} (Stetson \markcite{rmDAOPHOT} 1987) which can handle crowded fields by a simultaneous fit of profiles to a group of stars and gives superior photometric results, but is not designed for astrometry. Second, {\em SAAC} (Software for Analysing Astrometric CCD's) was used. SAAC was developed at Hamburg Observatory (Winter \markcite{rmW2} 1994) with various modifications and adaptions performed by N.Zacharias at USNO and performs 2--D fits on the star profiles with a choice of various models. Slightly saturated images of bright stars usually are of good astrometric quality as long as there is no bleeding into adjacent pixels. These images were kept, with caution, in the following reduction steps. Both program packages exclude significantly elongated images, mostly galaxies and various defects due to restrictions imposed on various image selection parameters. Positions of radio sources presented in this paper are based exclusively on the circular gaussian 2--D fit, which proved to be the best option for astrometry. For optical doubles the same fit model was used either with the cut--out or with a fit--up procedure (Schramm \markcite{rmSdiss} 1988) or both. In the cut--out option all pixels containing the companion image are excluded from the profile fit observation equations (interactively selected). In the fit--up procedure the background level is raised to a higher value up to the maximum count rate from the companion within the used window. Bright companion star images were first fitted by these methods exclucing the image of the etragalactic source. The thus obtained image profile was subtracted from the original pixel data and finally the position of the extragalactic source determined. The {\em allstar}(DAOPHOT) procedure was used for comparison in some cases. \subsection{Comparison of $x,y$ Data} Unique star numbers were assigned to all images of all stars of each field using our multi--step match program, based on position only. Coordinates were corrected for approximate third--order optical distortion, derived from a pilot investigation of a subset of the frames. Then all matches within a large search radius are recorded and the distribution of the coordinate differences of the matches in x and y are analyzed for a peak. Only the most likely identified matched images are used for a linear transformation using an unweighted least squares adjustment. Finally the transformed coordinates are compared to the reference coordinates and a position match with small tolerance is performed. Optionally, all images identified as multiple on a few arcseconds level are rejected except for manual 'fine tuning' for some radio source images. A transformation program was developed for comparing $x,y$ data of all overlapping CCD frames of the same field. Various mapping models are available and the residuals can be analyzed and displayed with already existing software from the photographic plate reduction package. The transformation program was used for comparing $x,y$ coordinates of the same CCD frame as obtained by different pixel fit algorithms, as well as to estimate field distortions and positional accuracy. Optionally, the obtained transformation parameters were applied to {\em combine} the $x,y$ data of all frames of each field into a superframe, from which spherical coordinates could be derived. \subsection{Combining $x,y$ Data} One option to derive positions from the combination of all CCD frames of a field is the traditional way of adjusting each individual CCD frame's $x,y$ data to spherical coordinates ($\alpha, \delta$) by means of reference stars and then combining these to obtain (weighted) mean positions. This option is not likely to give best results here. The number of available reference stars is usually small. Also, the deep exposure frames have overexposed images of the reference stars and an iterative process involving tertiary reference stars' $\alpha, \delta$ then is required, starting from the short exposure frames. Combining the $x,y$ data first is feasible without loss in accuracy due to the small field of view, the simple transformation geometry required here, and the large overlap in area of $\approx 40 \ to \ 100 \%$ for all our CCD frames of a field. Our software allows for a rigorous correction for refraction, but this effect was found to be negligible for our data. A significant third--order optical distortion term was removed before the weighted least squares adjustment with a linear transformation model for overlapping frames. Weights were obtained from the precision of the image profile fits, which strongly depend on the magnitude of the stars. In addition a constant variance per frame was added to account for atmospheric effects, scaled by the inverse exposure time of the CCD frames. This approach allows using simultaneously more reference stars in a larger field of view in cases when mosaic CCD frames are available, and at the same time combines all measurable images of long and short exposure frames into a common system. \subsection{Adjustment to Reference Star Positions} Our standard plate reduction software package, as part of HBAPP (Hamburg Block Adjustment Program Package, Zacharias \markcite{rmZdiss} 1987), was used for the unweighted least--squares adjustment of the $x,y$ data to the reference star positions. The precision of the secondary reference stars contributes the largest part of the errors in this adjustment. The differences in the precision of the CCD $x,y$ data, e.g. due to the dependence on magnitude or the atmospheric effects as a function of exposure time, are not relevant here. Experiments with weights obtained from the astrograph reductions did not reveal any significant improvement. Several pilot investigations were run to determine the most realistic mapping function of the telescopes (plate model). A linear plate model was finally adopted for routine processing with third--order optical distortion corrected prior to the adjustment. \subsection{External Comparisons} Field, as well as magnitude--dependent, systematic errors can be investigated externally by comparing $\alpha, \delta$ coordinates obtained from our CCD frames with positions obtained from a different telescope. For some fields ESO (European Southern Observatory) Schmidt plates, Lick Astrograph plates or prime focus plates from other telescopes are available. Due to the high precision of our CCD data a comparison with Schmidt plates only reveals systematic errors in those plates. Positions obtained from prime focus plates depend on the same secondary reference stars of that field already used for the CCD reductions. Thus the Lick Astrograph offers the most promising external comparison. Plate measuring is in progess and results will be published in an upcoming paper. Here only a comparison between different KPNO and CTIO runs will be made. A comparison of the optical positions with the quasi--error--free radio position can also reveal systematic errors. But our aim is to calibrate the optical data independently of that comparison in order to draw astrophysically significant conclusions. \section{Results} Because no photometric observations were obtained, all magnitudes in this paper are instrumental with the zeropoint adjusted to the mean photographic V magnitudes of the secondary reference stars. \subsection{Internal $x,y$ Precision} In this section we will compare the astrometric performance of various image profile fit models used to obtain $x,y$ coordinates from the pixel data of the CCD frames. Initial tests with the 1--dimensional center of mass algorithm showed significantly larger errors in positions as compared to other algorithms and thus was not further investigated. We compared the 1--dimensional gaussian fit (1DG) and the {\em allstar} point spread function fit (ALS) from the DAOPHOT/IRAF package, as well as the 2--dimensional circular gaussian fit (2DG) from our astrometric software package (SAAC). All comparisons were made only with stars appearing on all lists of $x,y$ coordinates obtained for the different profile fit models per frame, thus a unique limiting magnitude was used. All $x,y$ data were corrected for third--order optical distortion by applying the same value for the distortion coefficient (D3) and the location of the optical axis on the frame ($x_{0}, y_{0}$) to all frames prior to the transformation with a linear model. For routine reduction of the CCD frames in order to derive positions of extragalactic sources, the 2--dimensional circular gaussian (2DG) was adopted as our standard fit model because of its superior performance with respect to random, as well as systematic, errors as will be explained in the following two subsections. \subsubsection{KPNO 0.9 m Telescope} Table 3 shows some results for the KPNO field of the source 1656+053. The CCD frames with internal numbers 68 and 74 are long exposures (300 sec), while all others are short exposures (40 sec). Frames 68,74 and 73 are centered on the QSO radio source, while all others are part of a mosaic with offsets of 500" (735 pixel) in each coordinate. The seeing for all frames was $FWHM \approx 2.5$ pixel. Most of the $x,y$ data obtained from the 1DG fit show a significant magnitude--dependent error in both coordinates for the faint stars as compared to the $x,y$ positions from the same CCD frame obtained by the other fit models. An example is shown in Figure 1 a) for the x coordinate of frame 73 in the comparison of fit model 2DG - 1DG. The other coordinate, as well as the data from other frames, looks similar. The average position difference, as obtained by different profile fit models for the same pixel data, is on the order of 0.01 to 0.03 pixel, which is about 7 to 20 mas for the KPNO telescope. This is clearly a function of magnitude, increasing to the faint end. Figure 2 a) shows an example of the fit precision, $\sigma_{fit}$, as a function of magnitude for the 2DG fit of the pixel data obtained at KPNO from a 300--second exposure CCD frame of the field 0906+015. The saturation limit for this frame is at $\approx 13.0^{m}$. A best positional precision of 0.01 pixel is found for star images in the magnitude range from the saturation limit to 3 magnitudes below. For faint stars the fit precision decreases sharply. Outliers have been visually inspected on the CCD frame and only galaxies and close double stars have been found to cause significantly larger fit errors than the mean for that magnitude, providing a good method for detecting non--stellar images. Positions from multiple CCD frames of the same exposure time and the same field center were compared next. The standard error $\sigma_{xy}$ of such a frame--to--frame position transformation includes, besides the fit error $\sigma_{fit}$, also errors introduced by the atmospheric turbulence $\sigma_{atm}$. Figure 3 a) shows an example of $\sigma_{x}$ vs. magnitude for 2 frames from the field 0906+015 with a 300--second exposure time each. The limiting precision of $\approx 5 $ mas can entirely be accounted for by the turbulence in the atmosphere. According to Lindegren \markcite{rmLin} (1980) we have for our case $\sigma_{atm} \approx 10 $ mas. Compared to the previous figure either $\sigma_{fit}$ or $\sigma_{atm}$ or both are overestimated. An explanation for overestimating $\sigma_{fit}$ for bright stars is the difference of the observed image profile as compared to the assumed model. On the other hand, significant nightly variations of $\sigma_{atm}$ are also well known. Positions obtained for the brighter stars have been found to be less dependent on the profile fit algorithm used than those obtained for fainter stars. There was no systematic radial difference vs. radius found in any fit model comparisons of the same KPNO CCD frame. The transformation of $x,y$ coordinates of the frames with short exposure times show a larger sigma than those of long exposure times for all fit models (Table 3) because of the noise added by the atmosphere. In comparing the performance of the different fit models, the 2DG shows the smallest random errors for the KPNO frames. \subsubsection{CTIO 0.9 m Telescope} Table 4 shows some of the results for the CTIO telescope for the field 0646--306. Frames 53 and 54 are long exposures (600 sec, 300 sec), while all others are short exposures (40 sec). Frames 53, 54 and 57 are centered on the QSO, while all others are part of a mosaic with offsets of 240" = 600 pixel in both coordinates. Frame 53 has the poorest image quality as compared to the other frames. Contrary to the results obtained with the KPNO telescope, here the simple 1DG fit is in good agreement with the 2DG fit. An example is shown in Figure 1 b) for the difference in x--coordinates vs. magnitude for frame 57. No magnitude--dependent systematic errors were found in the test field. The CTIO telescope has better sampling and with FWHM $\approx 3.8 \ px$ in this test field there is a sufficient large number of pixels for both algorithms to determine consistent positions over a dynamic range of almost 8 magnitudes. With the CTIO data the DAOPHOT {\em allstar} algorithm show small magnitude--dependent systematic differences ($\approx 10 $ mas/mag) as compared to the 2DG and 1DG fit results as well as $\approx 20 $ mas systematic differences as a function of position in the field, when used with a single average point spread function (PSF) for the entire field of view. Figure 4 shows an example for the radial difference (2DG-ALS) vs. radius in frame 57. Clearly visible slightly elongated images at the edges of many CTIO frames require field--dependent PSF's to be used in DAOPHOT in order to obtain better results. The difficulty is how to relate these PSF's to each other astrometrically in order to get $x,y$ coordinates in a unique system for all stars in that frame on the 0.01 pixel level. In the frame--to--frame comparison again the 2DG shows the best results as judged from the standard deviations of the transformations of the $x,y$ data. The 1DG algorithm performs nearly as good as the 2DG in this respect, while the ALS is clearly inferior. For comparison with the KPNO results, Figure 2 b) shows a plot of the fit precision vs. magnitude. The limiting precision here is only $\approx $ 0.02 pixel. This can be explained by the poorer image quality of the CTIO as compared to the KPNO telescope, with variable deviations from the circular symmetric gaussian image profile depending on the location in the field. Expressed in arcseconds, both telescopes perform to about the same level of precision due to the better scale of the CTIO telescope. Figure 3 b) shows a frame--to--frame comparison, $\sigma_{x}$ vs. magnitude; both frames have been exposed for 600 seconds. Again a limit of $\approx 5 $ mas in precision is reached for bright stars in a {\em single} exposure. The atmosphere seems to be the limiting factor. \subsection{Basic Mapping Model} In this chapter the appropriate mapping model between the measured $x,y$ coordinates of the CCD frames and the corresponding standard coordinates ($\xi, \eta$) will be investigated. In order to allow for possible differences in scale and non--orthogonality of the axis a full linear transformation was adopted as our basic model. With orthogonal and non--orthogonal terms separated we have \[ \xi = a x + b y + c + e x + f y \] \[ \eta = -b x + a y + d + e y - f x \] \subsubsection{Optical Distortion Coefficient} A third--order optical distortion term (D3) was determined from $x,y$ data of mosaic frames. A conventional plate adjustment (CPA) of even a field with many ($\approx 20$) reference stars revealed no significant D3 term. The mean error on the D3 term is approximately $1.0 \times 10^{-9} \ "/"^{3}$. In order to obtain a reliable value for the D3 term a procedure similar to that of the AGK 2 catalog project (Schorr \& Kohlsch\"utter, \markcite{rmAGK2} 1951) was followed without the need for reference stars. Pairs of overlapping CCD frames with offsets in their centers in the order of half a field size have been transformed onto each other by extending the linear model with the appropriate D3 term \[ \Delta \xi = D3 \ (x_{1} r_{1}^{2} - x_{2} r_{2}^{2} ) \] \[ \Delta \eta = D3 \ (y_{1} r_{1}^{2} - y_{2} r_{2}^{2} ) \] with $x_{1},y_{1}$ and $x_{2}, y_{2}$ being the measured coordinates with respect to the center of distortion on frame 1 and frame 2 respecively and $ r_{i}^{2} = x_{i}^{2} + y_{i}^{2} \ , \ i=1,2 $. This algorithm assumes a common distortion term D3 for both frames, which is very realistic for frames taken shortly after each other with the same instrument and roughly the same location in the sky. This assumption was verified by comparing results for D3 obtained from various frame pairs. The mean values for D3 and their errors for both telescopes are given in Table 5 along with the maximal effect per coordinate on these 2K CCD's. The D3 term is highly significant and can be determined very precisely by this method. For convenience, conversion factors between the different units for quadratic and third--order terms in the CPA process are given in Table 6. \subsubsection{Optical Distortion Center} A significant offset of the center of distortion (optical axis) with respect to the geometric center of the CCD frame of $250 \pm 50$ pixel was found for the CTIO telescope in observing run 4. A similar offset of $\approx 100 \pm 70$ pixel in another direction was found for the same telescope in observing run 3, while no such offsets were found for the KPNO telescope (observing runs 1,2). Figure 5 a) shows a vector plot of average differences (run 4 - run 3) of $x,y$ data of field 0743-006 with optical distortion applied at the geometric frame centers prior to combining the $x,y$ data of frames for each run. Figure 5 b) shows the corresponding plot with optical distortion applied with respect to the optical axis as determined in a pilot investigation. In Figure 5 a) there is a systematic error of $\approx 70$ mas {\em at the frame center} which would have affected all source positions of that observing run. According to CTIO staff, such an offset of $\approx 200$ pixel is within the collimation tolerances of the instrument setup for each new observing run. \subsubsection{Tilt Terms} A difference in tangential points of two overlapping frames causes tilt terms (p,q) of the form (K\"onig, \markcite{rmK} 1933) \[ \Delta \xi = p x^{2} + q xy \] \[ \Delta \eta = p xy + q y^{2} \] Similar terms arise when individual frames are not perpendicular to the optical axis or the location of the tangential point is uncertain. A maximum difference in the location of tangential points for overlapping mosaic frames of 10 arcminutes was used here. This results in p,q terms as large as $ 1.4 \times 10^{-8} "/"^{2}$ (see Table 6). A maximum effect of $\Delta x \approx \Delta y \approx 0.01 \ pixel \ \le \ 7 $ mas is thus predicted for the edge of the field of view, which is totally negligible. Even assuming a tilt of the CCD plane with respect to the focal plane of the telescope (e.g. due to misalignment of the CCD camera) of $1^{\circ}$ results in a maximum effect of 0.06 pixel at the edge of the CCD frame, which would have little effect on the CPA results. As expected, no significant p,q terms were found neither in the $x,y$ transformation of overlapping plates, nor in the CPA of selected fields. A CPA with typical secondary reference stars is about a factor of 100 less sensitive to detect p,q terms as is the $x,y$ data transformation of overlapping frames. \subsection{External Calibration} Both instruments show field--dependent systematic errors of the order of $\approx 20 $ mas after applying the basic mapping model including third--order optical distortion (see e.g. Figure 5b). Unfortunately, presently no external calibration with respect to a precise reference star catalog can be made. Plates were taken at the Lick Astrograph with very small epoch difference (2 months) from some CCD observing runs. When Hipparcos results will become available, a position catalog to $\approx 30 $ mas precision for individual stars in an area of a few square degrees and down to 15th magnitude can be constructed for these external calibrations of our CCD data. A rigorous calibration of these FDP's (field distortion patterns) will then be possible, similar to the procedure used in photographic astrometry (Zacharias, \markcite{rmFDP} 1995). \subsection{Optical Positions of Reference Frame Sources} Here we present position results for 16 sources, selected as a representative subset of all optical counterparts on the current list of candidates ($> 400$ sources) for the extragalactic radio--optical reference frame. The sources were selected from all 4 CCD observing runs, sampling all areas in the sky as well as a wide range in magnitudes. Problematic cases, e.g. close doubles and sparce fields were prefered in order to challenge the reduction technique. Sources with multiple observations (different observing runs) as well as fields with more than one set of seconday reference stars were selected in order to obtain accuracy estimates by external comparisons. Table 7 gives a summary of the results. The positions are in the FK5/J2000 system as represented by the IRS catalog, and based on the 2--dimensional circular gaussian (2DG) fit model. The radio positions are taken from our radio reference frame (Johnston \markcite{rmRORF} et al. 1995). The rms difference (radio--optical) is $\approx 100$ mas for $\Delta \alpha \cos(\delta)$ and $\Delta \delta$. Compared to the expected internal errors this only shows the "wobbles" of the current optical system and gives no insight into the accuracy of the CCD observations. A reduction based on a Hipparcos intermediate solution was performed for the Hipparcos Working Group on Reference Link and shows an external error consistent with our error estimations. Detailed results will be published when the final Hipparcos catalog becomes available. \subsection{Remarks on Individual Sources} The following sources are optical doubles, most likely due to a foreground star. The images of both the extragalactic source and the companion are consistent with a stellar profile in all cases. No extended structure, e.g. of an underlying galaxy, was detected. All these sources are suitable for the radio--optical reference frame link, at least for now at the 30 mas level. 0153+744 is an optical double with a separation of $\approx 7 \ arcsec$ (10 pixel), which could be resolved without any problems. 0605-085 is an optical double with a flux of the companion 5 times {\em brighter} than the extragalactic source and a separation of 3 arcsec (4.5 px at KPNO, 7 px at CTIO). A subtraction of the image profile of the companion was required to obtain a position of the extragalactic source. 0607-157 is an optical double with a companion 2 times weaker than the extragalactic source with a separation of 4 arcsec (6 px at KPNO, 10 px at CTIO). The cut--out and fit--up procedures (see Section 4) allowed consistent image profile fits. 0743-006 is an optical double with a companion about a factor of 4 weaker than the extragalactic source and separated by 2.7 arcsec (6 px at CTIO). The fit--up and cut--out procedures with the 2DG fit gave consistent positions within 10 mas, while the {\em allstar} result was different by $\approx 30$ mas. Figure 6 shows a contour plot of this source, obtained from a 200--second exposure CTIO CCD frame in 1.5 arcsec seeing (3.7 pixel FWHM). 1800+440 is an optical double (3 arcsec) which was successfully fitted with the {\em allstar} algorithm as well as with the 2DG with use of cut--out and fit--up procedures. A contour plot of this source has been shown elsewhere (Zacharias et al. \markcite{rmZ166} 1995). The large number of optical doubles (5 out of 16 sources) presented here is not representative for all observations. These objects were selected for this pilot investigation. \subsection{Comparison of Multiple Data Sets} The sources 0336-019, 0605-085, 0607-157 and 0743-006 have been observed in more than one observing run. The mean quadratic difference in positions of the same source (see Table 7) obtained from different runs is $\approx 25 \ mas$, showing the high accuracy of the CCD observations. For 0743-006 there are 2 sets of secondary reference stars available from the Hamburg and Black Birch astrographs respectively. The agreement between the sets of secondary reference stars is on the same $\approx 25$ mas level, indicating the high accuracy of the secondary reference stars and the successful control of possible magnitude--dependent systematic errors in the astrograph fields. \section{Discussion} \subsection{Astrometric Properties of Both Telescopes} Based on the results of the previous section we estimate some individual random and systematic error contributions for observations made with both telescopes. The error of a single $x,y$ observation depends on the magnitude of the object. For faint objects the photon statistics limit the precision and the internal error obtained from the profile fit or the $x,y$ transformation from frame to frame is a good estimate of the accuracy of the position. For bright objects the random errors are as low as $\approx 0.01$ pixel and the systematic errors dominate. The astrometrically usable dynamic range depends on the pixel scale (positions for faint stars are getting worse when undersampled), as well as on the sky background level. Although most of our candidates are bright enough to be observed successfully within the full--Moon period, this has compromised the usable dynamic range, stressing the importance of the additional short exposure frames. Depending on the magnitude of the extragalactic reference link sources, a typical value for the random error of an $x,y$ coordinate of a single image on a CCD frame is $\sigma_{xy} \approx 15 $ mas, but the range was found to be as large as 5 to 31 mas (see Table 7). Systematic errors as a function of magnitude are expected to be $\le 10 $ mas over the magnitude range from secondary reference stars to extragalactic objects when using the appropriate profile fit model (2DG). Systematic errors as a function of location on the CCD frame are found on the $\approx 20 $ mas level, being larger far from the optical axis but negligible at the frame centers, where the image of the extragalactic object is usually located. These systematic errors will average out for different fields because of the different location of reference stars in each field. A more rigorous calibration of these FDP's is in progress which needs to be performed for each observing run separately, at least for the CTIO telescope. \subsection{Accuracy Estimate for the Link Procedure} In this section we will combine all estimates of individual error contributions for the optical observations of the reference frame link procedure based on counterparts of extragalactic radio sources. Special consideration is given to systematic errors in this multi--step procedure. We will refer to step 1 as the primary optical reference system. Options considered here are the IRS and the future Hipparcos and Tycho catalogs. By step 2 we denote the secondary reference stars which usually are obtained photographically by wide--field astrographs. An intermediate step 3 is sometimes taken with tertiary reference stars obtained either by Schmidt telesopes or CCD observations with wide fields. The last step is always the optical observation of the radio source counterpart itself, either photographically or with CCD. All error estimates given here are approximate. The aim is to identify the largest error contribution, to compare the performance of different options and to find the limits of this approach to the radio--optical reference frame link procedure. All formulae and values to follow are for one coordinate. As can be seen from the results in Table 7, the standard error of unit weight $\sigma_{CPA}$ of the adjustment of CCD $x,y$ to secondary reference star positions can be as low as $\approx 45$ mas, indicating a sub--micrometer accuracy from a single astrograph plate. Values for $\sigma_{CPA}$ increase with epoch difference of the secondary reference star observations and the CCD observations. This is due to the unknown proper motions in the secondary reference star data. No systematic corrections (e.g. galactic rotation) were applied here. \subsubsection{Algorithm} First we will define some quantities to be used in the link of step $i$ to step $i+1$. Let $n_{i}$ be the number of stars to be used as reference stars to link step $i$ to step $i+1$. The random error of such a link star in step $i$ we denote with $\sigma ran_{i}$, and $\sigma sys_{i+1}$ is the systematic error for the link of step $i$ to $i+1$. Similarly, $\sigma xy_{i+1}$ is the precision (random error) of a single $x,y$ observation for a link star in step $i+1$, and $m_{i+1}$ is the number of observations (exposures) for each link star in step $i+1$. The random error of a star position obtained in step $i+1$ then is approximately \begin{equation} \sigma ran_{i+1} \ = \ \frac{\sigma xy_{i+1}}{\sqrt{m_{i+1}}} \end{equation} Because there is only a limited number of link stars with associated errors between the two steps, the link of the system of step $i$ and $i+1$ can not be made error--free. The uncertainty in the zeropoint offset, $\sigma z_{i+1}$, between the coordinate systems in step $i$ and $i+1$ is approximately \begin{equation} \sigma z_{i+1} \ = \ \sqrt{\frac{\sigma^{2}ran_{i} \ + \ \sigma^{2}ran_{i+1}} {n_{i}}} \ = \ \frac{\sigma_{CPA}} {\sqrt{n_{i}}} \end{equation} Here $\sigma_{CPA}$ is the standard error of unit weight in the least--squares adjustment of {\em combined} $x,y$ data of step i+1 to reference star data of step i. But these formulae hold only for the central area of a frame (plate) and for a simple (linear) mapping model. In addition, a factor larger than one is required for other cases, and a rigorous derivation is given by Eichhorn \& Williams \markcite{rmEW} (1963). Our zeropoint offset can be considered as a special case of the error contribution due to error progapation of the plate constants to field star positions. Finally, the accuracy of a position of an extragalactic link source, $\sigma_{Q}$ , is approximately the rms sum of all zero point offsets from previous steps plus the systematic errors and the precision, $\sigma ran_{Q}$, of all optical observations of the source itself \begin{equation} \sigma_{Q} \ = \ \sqrt{\sigma^{2}ran_{Q} \ + \ \sum_{i=2}^{k} \sigma^{2}z_{i} \ + \ \sum_{i=2}^{k} \sigma^{2}sys_{i}} \end{equation} With systematic errors we mean here errors not averaging out with the number of stars used for the link of a single extragalactic source. It is assumed that such systematic errors (e.g. depending on magnitude for a particular plate) will be different for different extragalactic source fields and thus (at least partly) random when results for many sources are combined. Systematic errors inherent in this technique and not averaging out with different fields can't be investigated here. An external comparison with other methods for the extragalactic link procedure will be made in the future. \subsubsection{Accuracy Estimate for the Secondary Reference Star Positions} Here we will start out with 3 options for the primary optical reference system (step 1) and discuss 12 cases for determining secondary reference star positions (step 2). All cases are summarized in Table 8. The currently available IRS system has a density of $\approx 0.9 \ stars/degree^{2}$ and a precision of $\approx 200 $ mas for epochs of 1980 to 1994, where most of our data were taken. The usable field of views for the Hamburg Zone Astrograph (ZA), the Black Birch Astrometric Observatory (BBAO) astrograph and the Lick Astrograph are approximately 36, 25 and 9 $degree^{2}$ respectively. For the Hipparcos catalog we assume a mean $\sigma ran_{1} = 10 $ mas for the epoch range of our data. After the Tycho catalog is combined with the Astrographic Catalog (AC) data in order to obtain proper motions, we assume a mean $\sigma ran_{1} = 50 $ mas for that catalog at the epoch of our data. Systematic errors depending on magnitude are controlled with a diffraction grating at the astrographs. Preliminary results indicate magnitude terms on the order of $0 \ to \ 1 \ \mu m $ per 5 magnitudes, this is up to $20 $ mas/mag. The error on determining this term is about $ 2 $ mas/mag. This is a systematic error for a plate or field, which varies from field to field. All cases 4 are based on a $ 1 \ deg^{2}$ CCD frame. Cases 4b and 4c assume a mini--block adjustment of an area of $\approx 9 \ deg^{2}$. As can be seen from Table 8, a considerable improvement will be gained from the Hipparcos catalog as compared to the current IRS. A further improvement can be obtained by using the Tycho catalog, but this means a tremendous effort on plate measuring and the availability of the AC in order to derive good proper motions to be combined with the original Tycho observations. This is only worth the effort if the systematic errors can be controlled to this level. The CCD option is only competitive here when used with block adjustment techniques, at least in local fields. Because of the expected lower systematic errors, e.g. as a function of magnitude, this may be the way to go in the future. Such CCD observations could be based directly on Hipparcos stars with about the same precision as could be obtained from a Tycho--based solution, thus excluding possible systematic errors from the Tycho proper motions. \subsubsection{Position Accuracy of the Extragalactic Sources} With the algorithm as above and a typical internal precision of a CCD observation of $\sigma xy \approx 15 $ mas and $m = 2$ observations per source, we have a random error for the optical position of an extragalactic source of $\sigma ran_{Q} \approx 11 $ mas. Random errors from the CCD observations of the link stars (secondary reference stars) are even smaller due to their brightness, a typical value is $\sigma_{fit} \le 10 $ mas. The largest error contribution here is the influence due to the atmosphere in case of the short exposure frames. According to Lindegren \markcite{rmLin} (1980) this amounts to $\sigma xy_{3} \approx 30 $ mas for a single exposure; with $m_{3} \approx 4$ we have according to Eq. (1) $\sigma ran_{3} \approx 15 $ mas. Let us assume a random error from the astrograph observations of a link star (secondary reference star) of $\sigma ran_{2} \approx 70 $ mas. With $n_{2} \approx 15$ stars for that link between step 2 (secondary reference stars) and step 3 (CCD observations) we thus obtain according to Eq. (2) a zeropoint error of $\sigma z_{3} \approx 20 $ mas. This strongly depends on the number of stars used and the epoch difference between the CCD and astrograph observations. Individual results of the precision and accuracy properties of the CCD observations can be found in Table 7. Putting everything together, and assuming $\sigma sys_{2} \approx \sigma sys_{3} \approx 10 $ mas, we expect an accuracy of the optical position of an extragalactic object to be in the order of $\sigma_{Q} \approx 30 $ mas plus $\sigma z_{2}$ as discussed in the previous section, which is negligible in case of Hipparcos--catalog--based secondary reference stars ($\sigma z_{2} \approx 5 $ mas). When using the IRS catalog we have $\sigma z_{2} \approx 45 $ mas plus $\sigma sys_{1} \approx 100 $ mas. Individual estimates of $\sigma_{Q}$ (based on the Hipparcos catalog) are given in Table 7 for each object. Thus currently the largest error contribution comes from the primary system, the FK5, as represented by the IRS. With the use of the Hipparcos results the largest error contribution comes from the weak link of the secondary reference stars to the CCD observations due to the small field of view of the CCD's and the relatively poor limiting magnitude of the secondary reference stars. A large epoch difference between the secondary reference star and the QSO observations significantly increases the noise in this crucial step, regardless of any additional possible systematic errors indroduced by unknown proper motions. \subsection{Comparison to Other Investigations} Other major procedures for the position link of the radio and optical reference frame are the HST (Hubble Space Telescope) observations of selected pairs of Hipparcos stars and bright extragalactic candidates and the VLBI/VLA observations of Hipparcos radio stars. The HST observations are of higher internal precision than our observations but are not so numerous ($\approx 40$ pairs) and depend on the absolute calibration of the FGS fields. The radio star approch is very precise and direct but is based only on less than 10 objects which are not well distributed over the sky. Our approach contributes significantly to the link process and allows the important check to be made on possible systematic errors of the other methods. The Sloan Digital Sky Survey (SDSS) will be helpful in order to densify the grid of secondary (and tertiary) reference stars in the galactic north pole region. The positional accuracy of the optical counterparts of the extragalactic radio sources from SDSS will be inferior to our observations due to shorter integration time. \section{Conclusions} The feasibility of this approach to the radio--optical reference frame link has been proven here using wide field CCD observations with the KPNO and CTIO 0.9 m telescopes. The link to the primary reference star system, as represented by the Hipparcos catalog in the near future, is based entirely on photographic plates obtained with dedicated astrographs in both hemispheres. A deeper limiting magnitude and higher precision for the link stars is most important now, and CCD observations at the astrographs are in progress to provide more reference star positions. These observations will also provide an additional determination of possible magnitude--dependent systematic errors in the entire procedure. The 0.9 m telescope CCD observations have acquired a huge amount of high precision optical observations of extragalactic sources within a short period of time. The precision for a single long exposure is in the range of 5 to 31 mas (average $\approx 15$ mas) depending on the magnitude of the object. Field--dependent systematic errors exist on a 20 mas level, but they will be externally calibrated in the near future. A 20 mas precision level was reached previously with prime focus photography, but only for few objects per year at large telescopes. With 3 more CCD runs we hope to complete observations at the 0.9 m telescopes for about 400 sources which would allow a position tie to the Hipparcos system on the 1 mas level. In addition to the position information, a {\em structure} analysis of the optical counterparts is highly desirable. Because these objects have already been selected to be compact, the search for structure has to be made with much higher resolution than the 0.9 m ground--based telescopes can offer. Optical interferometry, adaptive optics or the HST are the only options at the moment. Currently our structure analysis is limited to identify suitable candidates for the link process, i.e. optical sources free of nearby disturbing foreground stars and galaxies. After a sufficiently rigid link between the radio and optical systems has been established, our observations will be used to identify outliers which will have astrophysical implications about the nature of these compact extragalactic objects. The biggest advantage of our approach to the extragalactic reference frame link is the large number of sources involved. Observations can easily be maintained in the future from ground--based telescopes, providing also an epoch difference large enough for a proper motion tie of the Hipparcos system within the next decade. With only minor improvements and more observations a much higher precision and accuracy can be reached by this technique in the near future. Position results of a large number of sources will be published after the Hipparcos catalog becomes available. No conclusions should be drawn from the positions published here based on the IRS system. When using a Hipparcos intermediate solution, as required for the Hipparcos Working Group on Reference Link, positional results with respect to the radio frame are in agreement with the error estimation given above. Ultimately a space mission like FAME (Johnston \markcite{rmFAME} 1995) or GAIA (Perryman \& van Leeuwen \markcite{rmGAIA} 1996) will provide optical positions for some (in case of FAME) or most (in case of GAIA) of these objects with an accuracy better than current VLBI radio observations. Until then we will hopefully have a much better understanding of the astrophysical and astrometric properties of these objects in order to be able to concentrate on the most suitable candidates for a reference frame. \acknowledgments Chr. de Vegt wishes to thank the Bundesministerium f\"ur Forschung und Technologie (BMFT) for financial support under Grant No. 50008810 (Hipparcos). We further thank J.L. Russell and M.I. Zacharias for assistance with observing, as well as J. M\"unkel for assistance with the astrographic plate measuring and reduction process. \newpage
\section*{} A method for making sure that the relativity effects are specified correctly (according to Einstein's General Relativity) can be described rather briefly. It agrees with Ashby's approach but omits all discussion of how, historically or logically, this viewpoint was developed. It also omits all the detailed calculations. It is merely a statement of principles. One first banishes the idea of an ``observer''. This idea aided Einstein in building special relativity but it is confusing and ambiguous in general relativity. Instead one divides the theoretical landscape into two categories. One category is the mathematical/conceptual model of whatever is happening that merits our attention. The other category is measuring instruments and the data tables they provide. For GPS the measuring instruments can be taken to be either ideal SI atomic clocks in trajectories determined by known forces, or else electromagnetic signals describing the state of the clock that radiates the signal. Each clock maintains its own proper time (but may convert this via software into other information when it transmits). We simplify to assume it transmits its own proper time without random or systematic errors, so that its increments $d\tau_T$ are simple physical data. Any other clock receiving these signals can record data tables showing the increments $d\tau_R$ in the SI proper time of the clock at the receiver corresponding to differences $d\tau_T$ in the proper times encoded in the signal it receives from some other identified transmitting clock. Once conventional zeros of time are identified for each clock, each transmitting and receiving pair produces a data table $\tau_R(\tau_T)$. These segments of data are to be reproduced by computations from the conceptual model, with any residuals understood on the basis of expected sources of noise and unmodelled phenomena. A user ``fix'' or relativistic ``event'' is the simultaneous reception of signals from four GPS satellites, or its equivalent from short extrapolations from nearly simultaneous signals. This user may not have a reliable clock but should be able to determine the time and position of the event from knowledge of the proper times encoded in the received signals, the identities of the transmitters, and the mathematical/conceptual model that defines the meaning of time and position for this purpose. System software aims to make the user calculations standard and practical, with many of the computational results encoded in the transmitted signals. What is the conceptual model? It is built from Einstein's General Relativity which asserts that spacetime is curved. This means that there is no precise intuitive significance for time and position. [Think of a Caesarian general hoping to locate an outpost. Would he understand that 600 miles North of Rome and 600 miles West could be a different spot depending on whether one measured North before West or visa versa?] But one can draw a spacetime map and give unambiguous interpretations. [On a Mercator projection of the Earth, one minute of latitude is one nautical mile everywhere, but the distance between minute tics varies over the map and must be taken into account when reading off both NS and EW distances.] There is no single best way to draw the spacetime map, but unambiguous choices can be made and communicated, as with the Mercator choice for describing the Earth. The conceptual model for a relativistic system is a spacetime map or diagram plus some rules for its interpretation. For GPS the attached Figure is a simplified version of the map. The real spacetime map is a computer program that assigns map locations $xyzt$ to a variety of events. In the Figure the $t$ time axis is vertical, and two of the three space $xyz$ axes are suggested horizontally. The wide center swath is the Earth which occupies the same location, centered on the central axis of the map, at all times. Marked on the surface of the Earth is a long spiral representing, e.g., a clock at USNO. The position of this clock as the Earth rotates is described by the coordinates of this curve on the (corresponding conceptual four dimensional) map, $x(t), y(t), z(t)$ where $xyzt$ are distances measured by a Euclidean ruler on the (conceptual four dimensional) graph paper parallel to its axes. The scale factors needed for interpreting this spacetime map are provided by the metric. In the map projection (coordinate system) from which the GPS model starts (an Earth Centered Inertial coordinate system, ECI) the metric is \begin{equation}\label{e-ECI} {d\tau}^2 = [1 + 2(V - \Phi_0)/c^2]dt^2 - [1 - 2V/c^2] (dx^2 + dy^2 + dz^2)/c^2 \quad . \end{equation} Here $V$ is the Newtonian gravitational potential of the Earth, approximately \begin{equation}\label{e-phi} V = - (GM/r) [1 - \half J_2 (R/r)^2 (3 \cos^2 \theta -1)] \quad . \end{equation} The constant $\Phi_0$ is chosen so that a standard SI clock ``on the geoid'' (e.g., USNO were it at sea level) would give, inserting its world line $x(t), y(t), z(t)$ into equation~(\ref{e-ECI}), just $d\tau = dt$ where $d\tau$ is the physical proper time reading of the clock. It is a theorem that if this choice is made for one clock on the geoid it applies to all. Equation~(\ref{e-ECI}) defines not only the gravitational field that is assumed, but also the coordinate system in which it is presented. There is no other source of information about the coordinates apart from the expression for the metric. It is also not possible to define the coordinate system unambiguously in any way that does not require a unique expression for the metric. In most cases where the coordinates are chosen for computational convenience, the expression for the metric is the most efficient way to communicate clearly the choice of coordinates that is being made. Mere words such as ``Earth Centered Inertial coordinates'' are ambiguous unless by convention they are understood to designate a particular expression for the metric, such as equation~(\ref{e-ECI}). Using equation~(\ref{e-ECI}) one can place tic marks along the world line of any clock to show changes in its proper time (which are to be physical changes directly displayed and transmitted by the clock). The computation is just to insert the clock trajectory $x(t), y(t), z(t)$ to find $d\tau$ from equation~(\ref{e-ECI}) as a thus specified multiple of $dt$. This applies both to Earth fixed clocks, to satellite clocks, and to clocks with any other motion $x(t), y(t), z(t)$ that has been incorporated in the map. The ``map'' here means a computer program that is designed to produce the trajectories $x(t), y(t), z(t)$ of each modelled object. The rules for drawing clock world lines or trajectories on the spacetime map (in the computer program) are simplest for dragfree satellites and for electromagnetic signals in vacuum. In these cases the world line must be a (timelike, resp.\ lightlike) geodesic of the metric~(\ref{e-ECI}), i.e., a solution of an ordinary differential equation constructed using the coefficients (scale factors) in equation~(\ref{e-ECI}). The electromagnetic signals have the special property that their trajectories also satisfy $d\tau^2 = 0$ in equation~(\ref{e-ECI}). By finding a lightlike geodesic that connects one tic mark $\tau_T$ on one clock world line to another mark $\tau_R$ on another clock, the map shows how one entry in the physical data table $\tau_R(\tau_T)$ is computed in the mathematical model. Once the observed data tables are being reproduced adequately in the mathematical model, its assignments of $xyzt$ coordinates to events identify the time and position of those events. In sum, the $txyz$ time and position values provided by GPS are not simple physical times and positions. Physical times and positions exist but, due to spacetime curvature, cannot be naturally associated with quadruples of numbers. Physical times and positions are identified on a spacetime map by their $xyzt$ map coordinates which depend on the ``projection'' (coordinate system) chosen in designing that particular map. The ECI map defined by equation~(\ref{e-ECI}) is the simplest to describe. More practical maps have been defined in which the space coordinates of geodetic benchmarks on Earth are nearly constant and change only due to tectonic and volcanic activity. To identify such an Earth fixed coordinate system one gives these coordinates as specified functions of those used in the ECI metric. This results in a metric expression different from equation~(\ref{e-ECI}) and allows results computed in the ECI coordinate system to be reported in the second coordinate system. \section*{Figure} This spacetime diagram shows the Earth, a fixed location (USNO) on the rotating Earth, a satellite orbiting the Earth, and an electromagnetic (EM) signal propagating from an event T on the satellite's world line to an event R on the USNO world line. Two of the three $xyz$ space axes are indicated. The $t$ time axis is at the center of the Earth. Any point on this diagram or map can be located by its $xyzt$ coordinates which are measured along the coordinate axes as conventional Cartesian coordinates for points (events) on this map. To deduce physical separations between (nearby) points on the map one must use equation~(\ref{e-ECI}) to convert the separations $dx\,dy\,dz\,dt$ read from the map into a physically measurable proper time interval $d\tau$. \section*{References} Neil Ashby, ``A tutorial on Relativistic Effects in the Global Positioning System'', NIST Contract No.40RANB9B8112, February 1990. Neil Ashby, ``Relativistic Effects in the Global Positioning System'', NIST Contract No.40RANB9B8112, August 1995. \end{document}
\section{Introduction}\indent The electrostatic potential of a homogeneously charged cube appears in theoretical studies of Wigner lattices \cite{Nijboer:88}. In computer simulations of ionic systems using minimum-image electrostatics, it determines the electrostatic self-interaction of ions \cite{Sloth:90,Sorensen:91,Hummer:93,Hummer:95:c,Figueirido:95}. In Ref.~\cite{Hummer:95:c}, Hummer {\em et al.} presented a simple calculation of the electrostatic potential at the center of a homogeneously charged cube. In this work, a closed form of the electrostatic potential will be determined for arbitrary positions. This analytic form can be used for the evaluation of lattice sums. It can also be applied as a correction when electrostatic potentials are calculated on a grid, assuming that the grid volumes are uniformly charged rather than carrying a point charge at the center. The analytic form of the potential will be compared with multipole expansions \cite{Nijboer:88,Durand:64:cube} \section{Calculation of the electrostatic potential of a cube}\indent The electrostatic potential $\phi_c$ of a cube $[-1/2,1/2]^3$ with charge density one will be calculated by integration. The potential at a point with Cartesian coordinates $(u,v,w)$ can be written as \begin{eqnarray} \phi_c(u,v,w) & = & \int_{-1/2}^{1/2} dx \int_{-1/2}^{1/2} dy \int_{-1/2}^{1/2} dz \left[ (x-u)^2 + (y-v)^2 + (z-w)^2 \right]^{-1/2}~, \label{eq:int} \end{eqnarray} where Gaussian units are used. $\phi_c$ can be rewritten as \begin{eqnarray} \int_{-1/2-u}^{1/2-u} dx \int_{-1/2-v}^{1/2-v} dy \int_{-1/2-w}^{1/2-w} dz \left( x^2 + y^2 + z^2 \right)^{-1/2}~. \end{eqnarray} Summation of the results of partial integration with respect to $x$, $y$, and $z$ yields a reduction to three two-dimensional integrals, \begin{eqnarray} \lefteqn{ 2 \phi_c(u,v,w) = \int_{-1/2-u}^{1/2-u} dx \int_{-1/2-v}^{1/2-v} dy \left[ z \left( x^2 + y^2 + z^2 \right)^{-1/2} \right]^{z=1/2-w}_{z=-1/2-w} }\nonumber\\ &&\mbox{+ cyclic permutations } (x,u;y,v;z,w)\rightarrow (y,v;z,w;x,u) \mbox{ and } (z,w;x,u;y,v)~. \end{eqnarray} The two-dimensional integrals can be further reduced using \begin{eqnarray} \int_{x_0}^{x_1} dx \int_{y_0}^{y_1} dy \left( x^2 + y^2 + z^2 \right)^{-1/2} & = & \int_{x_0}^{x_1} dx \; \frac{1}{2} \left[ \ln \frac{(x^2+y^2+z^2)^{1/2}+y}{(x^2+y^2+z^2)^{1/2}-y} \right] _{y_0}^{y_1}~, \end{eqnarray} where $x_0$, $x_1$, $y_0$, and $y_1$ are arbitrary integral boundaries. The remaining one-dimensional integrals can be calculated using partial integration and conventional substitution for algebraic integrands, \begin{eqnarray} \lefteqn{ \int dx \ln[(x^2+a^2)^{1/2}+b] = x \ln[ (x^2+a^2)^{1/2} + b ] -x}\nonumber\\ &&+ 2 | a^2 - b^2 |^{1/2} \;A\!\left[ \frac{x+(x^2+a^2)^{1/2}+b}{|a^2-b^2|^{1/2}}\right] + b \ln [ x + ( x^2 + a^2 )^{1/2} ]~, \end{eqnarray} where \begin{eqnarray} A(x) & = & \left\{ \begin{array}{lll} \arctan(x) & \mbox{for} & a^2 > b^2\\ \mbox{artanh}(x) & \mbox{for} & a^2 < b^2~. \end{array} \right. \end{eqnarray} Combining the previous results yields a closed form for the electrostatic potential of a unit cube: \begin{eqnarray} \lefteqn{\phi_c(u,v,w) = \frac{1}{2} \left\{ \frac{1}{2} \sum_{i=0}^{1} \sum_{j=0}^{1} \sum_{l=0}^{2} (-1)^{i+j} \; c_{i,l} \; c_{j,l+1} \right.} \nonumber\\ &&\times\ln\frac{ \left[\left( c_{i,l}^2 + c_{j,l+1}^2 + c_{1,l+2}^2 \right)^{1/2} + c_{1,l+2} \right]^3 \left[\left( c_{i,l}^2 + c_{j,l+1}^2 + c_{0,l+2}^2 \right)^{1/2} - c_{0,l+2} \right]}{ \left[\left( c_{i,l}^2 + c_{j,l+1}^2 + c_{1,l+2}^2 \right)^{1/2} - c_{1,l+2} \right] \left[\left( c_{i,l}^2 + c_{j,l+1}^2 + c_{0,l+2}^2 \right)^{1/2} + c_{0,l+2} \right]^3} \label{eq:c3}\\ &&\left. +\sum_{i=0}^{1} \sum_{j=0}^{1} \sum_{k=0}^{1} \sum_{l=0}^{2} (-1)^{i+j+k+1} \; c_{i,l}^2 \; \arctan\frac{ c_{i,l} \; c_{k,l+2}}{ c_{i,l}^2 + c_{j,l+2}^2 + c_{j,l+1} \left( c_{i,l}^2 + c_{j,l+1}^2 + c_{k,l+2}^2 \right)^{1/2} } \right\}~. \nonumber \end{eqnarray} The integration boundaries are defined as $c_{0,0} = - {1/2} - u$, $c_{1,0} = {1/2} - u$, $c_{0,1} = - {1/2} - v$, $c_{1,1} = {1/2} - v$, $c_{0,2} = - {1/2} - w$, and $c_{1,2} = {1/2} - w$. The values of $l+1$ and $l+2$ in Eq.~(\ref{eq:c3}) are defined modulo 3, {\em i.e.}, $c_{0,3} \equiv c_{0,0}$ etc.\ \ The $\arctan$ function to be used in Eq.~(\ref{eq:c3}) takes into account the sign of numerator and denominator and yields results between $-\pi$ and $\pi$ (``atan2'' in FORTRAN and C). An immediate consequence of Eq.~(\ref{eq:c3}) is the electrostatic potential at the center of a unit cube \begin{eqnarray} \phi_c(0,0,0) & = & 3 \ln \left( 3^{1/2} + 2 \right) - \frac{\pi}{2}~. \label{eq:phinull} \end{eqnarray} Previous calculations of $\phi_c(0,0,0)$ involved rather elaborate manipulations \cite{Nijboer:88,Sorensen:91}. \section{Calculation of the electrostatic potential of a square}\indent In two-dimensional electrostatics, the charge interaction (Green's function of the Laplacian) is given by $-\ln r$, where $r$ is the distance. The electrostatic potential $\phi_s$ of a square $[-1/2,1/2]^2$ with unit charge density will again be calculated by integration. $\phi_s$ is also the electrostatic potential of a square cylinder that is infinitely extended in $z$ direction. The potential at a point with Cartesian coordinates $(u,v)$ is written as \begin{eqnarray} \phi_s(u,v) & = & - \frac{1}{2} \int_{-1/2}^{1/2} dx \int_{-1/2}^{1/2} dy \ln \left[ (x-u)^2 + (y-v)^2 \right]~. \end{eqnarray} Elementary integration yields \begin{eqnarray} \phi_s(u,v) & = & - \frac{1}{2} \sum_{i=0}^{1} \sum_{j=0}^{1} \left[ x_i \; y_j \; \ln \left( x_i^2 + y_j^2 \right) - 3 \; x_i \; y_j + y_j^2 \; \arctan\frac{x_i}{y_j} + x_i^2 \; \arctan\frac{y_j}{x_i} \right]~, \nonumber\\ \label{eq:c2} \end{eqnarray} where $x_0 = -1/2 - u$, $x_1 = 1/2 - u$, $y_0 = -1/2 - v$, and $y_1 = 1/2 - v$. The appropriate $\arctan$ function to be used in Eq.~(\ref{eq:c2}), yields values between $-\pi/2$ and $\pi/2$ (``atan'' in FORTRAN and C). \section{Multipole expansion}\indent The electrostatic potential of a cube can be expanded in ``kubic'' harmonics, {\em i.e.}, harmonic functions with cubic symmetry \cite{Nijboer:88,Hummer:93,vdLage:47,Slattery:80,Adams:87}. For the exterior, one obtains \begin{eqnarray} \phi_c(\mbox{\bf r}) & = & \frac { 1 } { r } + C_4 \; K_4(\mbox{\bf r}) \; r^{-9} + C_6 \; K_6(\mbox{\bf r}) \; r^{-13} + \cdots ~, \label{eq:mout} \end{eqnarray} where $\mbox{\bf r}=(u,v,w)$, $r=|\mbox{\bf r}|$. With $T_n=u^n+v^n+w^n$, the kubic harmonics of order 4 and 6 can be written as \cite{Adams:87} \begin{eqnarray} K_4(\mbox{\bf r}) & = & T_4 - \frac{3}{5} \; r^4 \\ K_6(\mbox{\bf r}) & = & T_6 - \frac{15}{11} \; T_4 \; r^2 + \frac{30}{77} \; r^6~. \end{eqnarray} For this form, the expansion coefficients are $C_4=-7/192$ and $C_6=11/192$ \cite{Nijboer:88,Durand:64:cube}. For the interior of the cube, we derive the multipole-expansion coefficients of order 2, 4, and 6 from a direct Taylor expansion in $x$ direction. The angular dependence can then be inferred by cubic symmetry.\footnote{Some higher-order kubic harmonics are degenerate \cite{Slattery:80}, requiring two independent expansion directions to get the correct angular dependence.} The electrostatic potential on the $x$ axis can be expressed as \begin{eqnarray} \phi_c(u,0,0) & = & \int_{-1/2-u}^{1/2-u} dx \; f(x)~, \end{eqnarray} where \begin{eqnarray} f(x) & = & \int_{-1/2}^{1/2} dy \int_{-1/2}^{1/2} dz \left( x^2 + y^2 + z^2 \right)^{-1/2}\nonumber\\ & = & 2 \ln \frac{(4x^2+2)^{1/2}+1}{(4x^2+2)^{1/2}-1} - 2\;x\; \arctan\frac{4x(4x^2+2)^{1/2}}{16x^4+8x^2-1}~. \end{eqnarray} Taylor expansion of $\phi_c(u,0,0)$ around $u=0$ yields the expansion \begin{eqnarray} \phi_c(\mbox{\bf r}) & = & 3 \ln \left( 3^{1/2} + 2 \right) - \frac{\pi}{2} -\frac{2\pi}{3} \; r^2 - \frac{40}{243^{1/2}} K_4(\mbox{\bf r}) - \frac{308}{19683^{1/2}} K_6(\mbox{\bf r}) + \cdots~. \label{eq:min} \end{eqnarray} Figure~\ref{fig:axis} shows the electrostatic potential $\phi_c$ along the directions $(u,0,0)$, $(u,u,0)$, and $(u,u,u)$ calculated from the exact result Eq.~(\ref{eq:c3}) and the expansions Eq.~(\ref{eq:mout}) and (\ref{eq:min}), both including terms up to $K_4$. The expansions show the largest disagreement near the surface of the cube ($u=1/2$) where they start to diverge. Otherwise, they closely reproduce the exact potential [Eq.~(\ref{eq:mout}) for $r\rightarrow\infty$ and Eq.~(\ref{eq:min}) for $r\rightarrow 0$]. The divergent behavior reflects an inherent problem of the near- and far-field expansions. By construction, the Laplacians of Eqs.~(\ref{eq:mout}) and (\ref{eq:min}) are a delta function at $r=0$ and a constant $-4\pi$, respectively, independent of the order of the expansions. The former corresponds to a unit point charge and is correct only outside the cube; the latter corresponds to a homogeneous charge density and is correct only inside the cube. \section{Conclusion}\indent Nijboer and Ruijgrok \cite{Nijboer:88} analyzed the difference between the energy per particle in a Wigner lattice and the energy of a point charge in the field of the other charges. These authors studied an infinite replication of neutral cubes consisting of a unit point charge at the center and a compensating background. A reduction of the electrostatic potential $\phi_c$ of a homogeneously charged cube to a one-dimensional integral resulted in \begin{eqnarray} \phi_c(u,v,w) & = & \frac{\pi}{8} \int_{0}^\infty dt \; t^{-2} \; \frac{\partial}{\partial t} \left[ h(u,t) h(v,t) h(w,t) \right]~, \label{eq:NR} \end{eqnarray} where \begin{eqnarray} h(x,t) & = & \mbox{erf}\left[\left(x+\frac{1}{2}\right)t\right] - \mbox{erf}\left[\left(x-\frac{1}{2}\right)t\right] \end{eqnarray} and $\mbox{erf}$ is the error function.\footnote{Eq.~(2.6) of Ref.~\cite{Nijboer:88} is missing a factor $\pi$ on the right-hand side. Eq.~(2.8) of Ref.~\cite{Nijboer:88} has the correct pre-factor.} The solution of the one-dimensional integral in Eq.~(\ref{eq:NR}) would give the closed form Eq.~(\ref{eq:c3}) of this work. Eq.~(\ref{eq:c3}), numerical integration of Eq.~(\ref{eq:NR}), and direct Monte Carlo integration of Eq.~(\ref{eq:int}) were compared for a few hundred points and gave identical results within the error margins of the numerical integration in Eq.~(\ref{eq:NR}) and the statistical errors of the Monte Carlo procedure. Eq.~(\ref{eq:c3}) has the advantage of being analytical. It can be evaluated fast and with arbitrary precision on the computer. The electrostatic potential $\phi_c(0,0,0)$ at the center of the cube as listed in Eq.~(\ref{eq:phinull}) can be used to correct effectively for finite-size effects in computer simulations of ionic systems under periodic boundary conditions, when minimum-image electrostatics is used \cite{Allen:87}. An example is the calculation of single-ion chemical potentials \cite{Sloth:90,Sorensen:91,Hummer:93,Hummer:95:c,Figueirido:95}, where the electrostatic energy of an excess ion has to be calculated. The system-size dependence is greatly reduced if the excess charge is compensated with a homogeneous background. The electrostatic energy $u$ of the excess charge $q$ at ${\bf r}=0$ is then the sum of the interactions with the other charges $q_i$ at $\mbox{\bf r}_i$ and with the background, \begin{eqnarray} u & = & q \sum_{i=1}^{N} q_i / r_i + q^2 \phi_c(0,0,0)/L~, \end{eqnarray} where a cubical box of length $L$ is used. Another application is the calculation of electrostatic potentials when charges are given on a grid, for instance, when ionic density distributions are known \cite{Klement:91}. Usually, the grid charges are assumed to be point charges. In an improved description, the charges are smeared out over the grid cells. The electrostatic potentials can then be calculated using Eq.~(\ref{eq:c3}) or the multipole expansions Eq.~(\ref{eq:mout}) and (\ref{eq:min}). This eliminates the singularities in the electrostatic potential and gives a more accurate description near local charge concentrations.
\section{Introduction} \setcounter{equation}{0} A lot of progress has been made in the last few years in the understanding of S-duality as a symmetry of four dimensional gauge theories. The conjecture of Montonen and Olive \cite{mo} that $N=4$ supersymmetric Yang-Mills theories were invariant under strong-weak coupling with the exchange of the gauge group by its dual was tested in \cite{vw}, were it was shown that in fact the partition function transformed as a modular form. Some progress has been also made for $N=2$ and $N=1$ supersymmetric Yang-Mills theories \cite{sw,seiberg}. However a path integral derivation of S-duality is in general still unknown. In \cite{witten} Witten showed that S-duality in four dimensional abelian gauge theories \cite{cr,cardy,shw} can be implemented at the level of the path integral in a very similar way to T-duality in non-linear sigma models in String Theory \cite{reviews}. The idea is to consider a global isometry of the Lagrangian which can be expressed as translations of a given coordinate (the adapted coordinate), gauge this isometry by introducing a fake gauge field and impose the constraint that the curvature tensor associated to this gauge field is zero so that the gauge field is non-propagating. Integrating the Lagrange multiplier and fixing the gauge field to zero the original theory is recovered and integrating the gauge field and fixing the adapted coordinate to zero the new dual theory is obtained. In the case of T-duality the initial variables are 0-forms and the global isometry that is gauged is $\theta\rightarrow \theta+\epsilon$ where $\theta$ is the adapted coordinate. In the case of abelian gauge theories the initial variables are 1-forms and the isometry which is gauged is $A\rightarrow A+\epsilon$ where now the $\epsilon$ parameter is a 1-form. Then the gauge field which has to be introduced is a 2-form and its field strength a 3-form. In 4 dimensions the Lagrange multiplier imposing that the field strength vanishes is a 1-form, like the original gauge field, and the dual theory is expressed also in terms of 1-forms. Also for this non-supersymmetric case the partition function transforms as a modular function with a modular weight proportional to the Euler characteristic and the signature of the manifold \cite{witten,verlinde}. Given the analogy with T-duality a canonical transformation must be beyond this path integral manipulation, since this is the case in T-duality \cite{venezia,aagl2,la}. In section 2 we present the explicit generating functional producing this transformation and show that it is the generalization of the functional in 2-dimensional non-linear sigma models to 4 dimensions and 1-forms. Under this transformation electric and magnetic degrees of freedom get interchanged (with the minus relative sign) as shown in abelian lattice gauge theories in \cite{cardy}. The canonical transformation approach is the simplest in order to obtain the dual theory, also in this case in which in the Hamiltonian formulation one has to be careful with the constraints. It is easy to show that both the initial and the dual theory are defined in the same subspace of the phase space after the canonical transformation is performed. We show that in phase space the partition functions of the initial and dual theories coincide and that only after integrating out the momenta degrees of freedom the modular anomaly \cite{witten,verlinde} appears. The same canonical transformation applied to the non-abelian case seems to relate Yang-Mills theories with inverted couplings. However a careful analysis of the constraints points out that this is not the case. The dual theory is in fact of Freedman-Townsend's type \cite{ft}, i.e. it is expressed as a function of arbitrary 2-forms which are not derived from a vector potential. We show this in section 3. The results presented in section 2 can be easily generalized to the case of $d$ dimensional abelian gauge theories of $p$-forms, as it is explained in section 4. The modular anomaly in the transformation of the partition function is obtained. The implementation at the level of the path integral using a coset construction was presented in \cite{barbon}. \section{The abelian case} \setcounter{equation}{0} In this section we construct the explicit canonical transformation which produces the change \begin{equation} \label{2uno} \tau\rightarrow -1/\tau \end{equation} with $\tau=\theta/2\pi+4\pi i/g^2$, for $U(1)$ four dimensional euclidean gauge theories. Let us consider the Lagrangian \begin{eqnarray} \label{2dos} L&=&\frac{1}{8\pi}(\frac{4\pi}{g^2}F_{mn}F^{mn}+\frac{i\theta}{4\pi} \epsilon_{mnpq}F^{mn}F^{pq}) \nonumber\\ &=&\frac{i}{8\pi}({\bar \tau} F^+_{mn}F^{+mn} -\tau F^-_{mn}F^{-mn}) \end{eqnarray} where \begin{eqnarray} \label{2tres} F^+_{mn}&=&\frac12 (F_{mn}+\,^*F_{mn})=\frac12 (F_{mn}+\frac12 \epsilon_{mnpq}F^{pq}),\nonumber\\ F^-_{mn}&=&\frac12 (F_{mn}-\,^*F_{mn})=\frac12 (F_{mn}-\frac12 \epsilon_{mnpq}F^{pq}) \end{eqnarray} and $F_{mn}=\partial_m A_n-\partial_n A_m$. It was shown in \cite{witten} that the transformation (\ref{2uno}) could be derived at the level of the path integral by the usual Rocek and Verlinde's procedure \cite{rv} one follows to construct abelian T-duals of two dimensional sigma models in String Theory. In this case given a global abelian continuous isometry of the sigma model one can turn it local by introducing a fake gauge field in the Lagrangian by minimal coupling and imposing the constraint that this gauge field is non-dynamical. Solving this constraint and fixing the gauge field to be zero one recovers the original theory. If instead the gauge field is integrated and the gauge is fixed in the original variables a sigma model written in terms of the Lagrange multiplier introduced to impose the constraint is obtained. This is the dual sigma model. In \cite{witten} the same construction is applied to obtain the dual of the four dimensional abelian gauge theory. The global continuous abelian isometry in this theory is \begin{equation} \label{2cuatro} A\rightarrow A+\epsilon \end{equation} where now the isometry parameter is a 1-form. This global isometry can be gauged by introducing a gauge field $G$, 2-form, which is imposed to be non-dynamical with the term \begin{equation} \label{2cinco} \int_M d^4 x {\tilde A} dG \end{equation} where the Lagrange multiplier ${\tilde A}$ is a 1-form. Integrating ${\tilde A}$ the constraint $dG=0$ is obtained, ie. $G$ pure gauge, and we can recover (\ref{2dos}) by either fixing $A=0$ or $G=0$. On the other hand by integrating out $G$ and then fixing $A=0$ the following Lagrangian is gotten: \begin{equation} \label{2seis} {\tilde L}=\frac{i}{8\pi}(-\frac{1}{{\bar \tau}} {\tilde F}^+_{mn} {\tilde F}^{+mn} +\frac{1}{\tau} {\tilde F}^-_{mn}{\tilde F}^{-mn}) \end{equation} with ${\tilde F}^{\pm}$ the self- and antiself-dual components of ${\tilde F}_{mn}\equiv\partial_m {\tilde A}_n-\partial_n {\tilde A}_m$. This is the S-dual of the initial electromagnetic theory since in the particular case $\theta=0$ it corresponds to the inversion of the coupling constant $g$. In this procedure we have made an integration by parts in the Lagrange multipliers term and neglected a total derivative\footnote{This term is seen in the gauge $A=0$.}. However this total derivative contains some information, in particular it implies that the initial and dual Lagrangians are equal up to a total time derivative, exactly what happens when two theories are related by a canonical transformation. To be more precise, the generating functional of a canonical transformation from $\{q^i,p_i\}$ to $\{Q^i,P_i\}$ is such that \begin{equation} \label{2siete} p_i \dot{q^i}-H(q^i,p_i)=P_i \dot{Q^i}-\tilde{H}(Q^i,P_i)+ \frac{d{\cal F}}{dt}. \end{equation} If ${\cal F}$ is a type I generating functional (depending only on coordinates) $H=\tilde{H}$ if and only if\footnote{We assume ${\cal F}$ does not depend explicitly on time.} \begin{eqnarray} \label{2ocho} &&\frac{\partial {\cal F}}{\partial q^i}=p_i \nonumber\\ &&\frac{\partial {\cal F}}{\partial Q^i}=-P_i \end{eqnarray} Under duality\footnote{We have dropped the global $i/8\pi$ factor. It will then appear when exponentiating these quantities.}: \begin{equation} \label{2nueve} {\tilde L}({\tilde A})=L(A)+d{\tilde A}\wedge dA \end{equation} which implies\footnote{Our convention for the product of forms is: ${\tilde F}\wedge F=\epsilon^{mnpq}{\tilde F}_{mn} F_{pq}$.} \begin{equation} \label{2diez} \epsilon^{mnpq} (\partial_m {\tilde A}_n-\partial_n {\tilde A}_m) (\partial_p A_q-\partial_q A_p)=-(\frac{\delta {\cal F}}{\delta {\tilde A}_m} \dot{{\tilde A}}_m +\frac{\delta {\cal F}}{\delta A_m}\dot{A}_m) \end{equation} This produces the canonical transformation \begin{eqnarray} \label{2once} &&{\Pi}^\alpha=\frac{\delta {\cal F}}{\delta A_\alpha}= -4\,^*{\tilde F}^{0\alpha} ,\qquad \Pi^0=0, \nonumber\\ &&{\tilde \Pi}^\alpha=-\frac{\delta {\cal F}}{\delta {\tilde A}_\alpha} =4\,^*F^{0\alpha}, \qquad {\tilde \Pi}^0=0 \end{eqnarray} plus a constraint \begin{equation} \label{2doce} \Pi^\alpha\partial_\alpha A_0={\tilde \Pi}^\alpha\partial_\alpha {\tilde A}_0, \end{equation} where greek indices run over spatial coordinates. The generating functional producing this canonical transformation is \begin{equation} \label{2trece} {\cal F}=-2\int_{M, t fixed} d^3 x ({\tilde A}_\alpha\,^*F^{0\alpha}+ A_\alpha\,^*{\tilde F}^{0\alpha})=-\frac12 \int_M d^4 x {\tilde F}\wedge F. \end{equation} This is the result one would expect a priori from what is known in two-dimensional sigma-models, where the generating functional is given in terms of the adapted coordinate to the isometry $\theta$ and the Lagrange multiplier ${\tilde \theta}$ by \cite{venezia,aagl2} \begin{equation} \label{2catorce} {\cal F}=-\frac12 \int_{M_2} d{\tilde \theta}\wedge d\theta. \end{equation} The Hamiltonian associated to (\ref{2dos}) is given by: \begin{equation} \label{2quince} H=\frac{1}{4(\bar{\tau}-\tau)}\Pi_\alpha \Pi^\alpha+\partial_\alpha A_0 \Pi^\alpha-\frac{\bar{\tau}+\tau}{\bar{\tau}-\tau}\Pi_\alpha \,^*F^{0\alpha} +\frac{4\bar{\tau}\tau}{\bar{\tau}-\tau}\,^*F^{0\alpha} \,^*F_{0\alpha} \end{equation} plus the constraints \begin{equation} \label{2dieciseis} \Pi_0=0,\qquad \partial_\alpha \Pi^\alpha=0, \end{equation} where \begin{equation} \label{2dieciseisbis} \Pi^\alpha= 4{\bar \tau} F^{+0\alpha}-4\tau F^{-0\alpha}. \end{equation} $\Pi_0$ is a primary constraint and $\partial_\alpha \Pi^\alpha=0$ is the secondary constraint emerging from the equation of motion for $\Pi_0$. They imply that the theory is defined in the reduced phase space given by $\Pi_0=0$, $\partial_\alpha \Pi^\alpha=0$. These constraints are also satisfied in the dual theory, since they are obtained directly from the canonical transformation. Then the dual theory is defined in the same reduced phase space than the original one. The relation (\ref{2doce}) is trivial in this subspace. However we need to consider it in order to recover the dual Lagrangian from the canonically transformed Hamiltonian, since for that we need the naive Hamiltonian without taking into account the constraints. Our purpose is to show that the canonically transformed Lagrangian is the dual Lagrangian and for that we do not need to study in detail the way the theory gets defined in the Hamiltonian formalism \cite{Ramond}, it is enough to show that both the initial and dual theories are defined in the same reduced phase space. The canonically transformed Hamiltonian reads: \begin{equation} \label{2diecisiete} {\tilde H}=\frac14 \frac{\bar{\tau}\tau}{\bar{\tau}-\tau} {\tilde \Pi}_\alpha {\tilde \Pi}^\alpha+\partial_\alpha {\tilde A}_0 {\tilde \Pi}^\alpha+ \frac{\bar\tau +\tau}{\bar{\tau}-\tau}{\tilde \Pi}_\alpha\,^*{\tilde F}^{0\alpha}+ \frac{4}{\bar{\tau}-\tau}\,^*{\tilde F}_{0\alpha} \,^*{\tilde F}^{0\alpha}. \end{equation} The corresponding Lagrangian is given by the dual Lagrangian (\ref{2seis}). Recall that (\ref{2once}): \begin{eqnarray} \label{217a} &&\Pi^\alpha=-4\,^*{\tilde F}^{0\alpha}, \nonumber\\ &&{\tilde \Pi}^\alpha=4\,^*F^{0\alpha} \nonumber \end{eqnarray} corresponds to the usual interchange between electric and magnetic degrees of freedom when there is no $\theta$-term. Some useful information can be obtained within this approach. The generating functional (\ref{2trece}) is linear in both the original and dual variables. Then the following relation holds: \begin{equation} \label{217b} H e^{\frac{i{\cal F}}{8\pi}}={\tilde H}e^{\frac{i{\cal F}}{8\pi}} \end{equation} which implies: \begin{equation} \label{217c} \psi_k[{\tilde A}]=N(k)\int {\cal D}A(x^\alpha) e^{\frac{i}{8\pi}{\cal F}[{\tilde A},A(x^\alpha)]} \phi_k[A(x^\alpha)] \end{equation} with $\phi_k[A]$ and $\psi_k[{\tilde A}]$ eigenfunctions of the initial and dual Hamiltonians respectively with the same eigenvalue and $N(k)$ a normalization factor \cite{ghandour}. {}From this relation global properties can be easily worked out. The Dirac quantization condition: \begin{equation} \label{217d} \int_{\Sigma}F=2\pi n,\quad n\in Z, \end{equation} for $\Sigma$ any closed two-surface in the manifold, implies for ${\tilde F}$: \begin{equation} \label{217e} \int_{\Sigma}{\tilde F}=2\pi m,\quad m\in Z \end{equation} and ${\tilde F}$ must live in the dual lattice. Also from (\ref{217c}) the transformation applies to any four dimensional manifold $M$ since $\phi_k[A]$ can be the result of integrating the theory in an arbitrary manifold with boundary. We can obtain in phase space the modular anomaly emerging in the transformation of the partition function \cite{witten,verlinde}. The argument goes as follows. In phase space the partition function is given by\footnote{In order to have a well-defined partition function we have to fix the gauge symmetry. The following arguments are in this sense formal.}: \begin{equation} \label{2dieciocho} Z_{ps}=\int {\cal D}A_\alpha {\cal D}\Pi^\alpha e^{- \frac{i}{8\pi}\int d^4x (\dot{A}_\alpha \Pi^\alpha-H)} \end{equation} Under (\ref{2once}) \begin{equation} \label{218b} {\cal D}A_\alpha {\cal D}\Pi^\alpha={\cal D}{\tilde A}_\alpha {\cal D} {\tilde \Pi}^\alpha. \end{equation} Then the dual phase space partition function is given by: \begin{equation} \label{2diecinueve} {\tilde Z}_{ps}=\int {\cal D}{\tilde A}_\alpha {\cal D}{\tilde \Pi}^\alpha e^{-\frac{i}{8\pi}\int d^4x (\dot{{\tilde A}}_\alpha {\tilde \Pi}^\alpha-{\tilde H})}=Z_{ps} \end{equation} showing that in phase space the partition function is invariant under duality. Integration on momenta in (\ref{2dieciocho}) gives: \begin{equation} \label{2veinte} Z_{ps}=\int {\cal D}A_\alpha ({\rm Im} \tau)^{B_2/2} e^{-\int d^4x L} \end{equation} with $L$ given by (\ref{2dos}). The factor $({\rm Im} \tau)^{B_2/2}$ in the measure is the regularized $({\rm det\,Im}\tau)^{1/2}$ coming from the gaussian integration over the momenta. $B_2$ is the dimension of the space of 2-forms in the four dimensional manifold $M$ (regularized on a lattice) and emerges because the momenta are 2-forms. The same calculation in the dual phase space partition function gives: \begin{equation} \label{2veintiuno} {\tilde Z}_{ps}=\int {\cal D}{\tilde A}_\alpha ({\rm det} ({\rm Im} -\frac{1}{\tau}))^{1/2} e^{-\int d^4x {\tilde L}} \end{equation} with ${\tilde L}$ given by (\ref{2seis}). We regularize the factor \begin{equation} \label{2veintidos} ({\rm det}({\rm Im} -\frac{1}{\tau}))^{1/2}=({\rm det}({\rm Im} \tau /(\tau {\bar\tau})))^{1/2} \end{equation} by \begin{equation} \label{2veintitres} ({\rm Im} \tau)^{B_2/2}\bar\tau^{-B_2^+/2}\tau^{-B_2^-/2} \end{equation} where $B_2^+$ and $B_2^-$ are respectively the dimensions of the spaces of self-dual and anti-self-dual 2-forms. In configuration space the partition function is defined by \cite{witten}: \begin{equation} \label{2veinticuatro} Z=({\rm Im} \tau)^{(B_1-B_0)/2}\int {\cal D}A_\alpha e^{-S}= ({\rm Im} \tau)^{(B_1-B_0-B_2)/2} Z_{ps} \end{equation} and in the dual model \begin{equation} \label{2veinticinco} {\tilde Z}=(\frac{{\rm Im} \tau}{\tau \bar\tau})^{(B_1-B_0)/2} \int {\cal D} {\tilde A}_\alpha e^{-{\tilde S}}. \end{equation} {}From $Z_{ps}={\tilde Z}_{ps}$ we arrive to \begin{equation} \label{2veintiseis} Z=\tau^{-(\chi-\sigma)/4}{\bar\tau}^{-(\chi+\sigma)/4} {\tilde Z} \end{equation} where $\chi=2(B_0-B_1)+B_2$ is the Euler number (the regularization is such that $B_p=B_{4-p}$) and $\sigma=B_2^+-B_2^-$ is the signature of the manifold. This is the modular factor appearing in \cite{witten,verlinde}. In phase space the partition function is simply defined as the integration over coordinates and momenta and it transforms as a scalar with modular weight equal to zero. Is only when going to the configuration space that the integrations over the momenta produce some determinants which after being regularized yield the modular factor found in \cite{witten,verlinde}. A very similar argument should apply to the transformation of the dilaton in two-dimensional non-linear sigma-models. \section{The non-abelian case} \setcounter{equation}{0} The canonical transformation approach can be generalized to the case of non-abelian gauge theories with compact group $G$. The initial Lagrangian is given by: \begin{eqnarray} \label{4uno} L&=&\frac{1}{8\pi}(\frac{4\pi}{g^2}F^{(a)}_{mn} F^{(a)mn}+ \frac{i\theta}{4\pi}\epsilon^{mnpq}F^{(a)}_{mn} F^{(a)}_{pq}) \nonumber\\ &=&\frac{i}{8\pi}({\bar \tau} F^{(a)+}_{mn} F^{(a)+ mn}-\tau F^{(a)-}_{mn}F^{(a)- mn}) \end{eqnarray} where $F=dA-A\wedge A$ and we have chosen $Tr(T^a T^b)=\delta^{ab}$ ($T^a$ are the generators of the Lie algebra). The conjugate momenta and the Hamiltonian are: \begin{eqnarray} \label{i1} &&\Pi^{a\alpha}= 2({\bar \tau}-\tau)F^{(a)0\alpha}+2({\bar \tau}+\tau)\, ^*F^{(a)0\alpha} \nonumber\\ &&\Pi^{a0}=0 \end{eqnarray} \begin{equation} \label{i2} H=\frac14 \frac{1}{{\bar \tau}-\tau}\Pi^a_\alpha \Pi^{a\alpha}+ (\partial_\alpha A^a_0+f_{abc}A^b_0A^c_\alpha)\Pi^{a\alpha}- \frac{{\bar \tau}+\tau}{{\bar \tau}-\tau}\Pi^{a\alpha}\, ^*F^{(a)}_{0\alpha}+\frac{4{\bar \tau}\tau}{{\bar \tau}-\tau} \,^*F^{(a)}_{0\alpha}\,^*F^{(a)0\alpha}, \end{equation} with $f_{abc}$ the structure constants of the Lie algebra. The equations of motion of the primary constraints $\Pi^{a0}=0$ imply: \begin{equation} \label{41bf} \partial_\alpha \Pi^{a\alpha}-f_{abc}A^b_\alpha\Pi^{c\alpha}=0, \end{equation} so that we can ignore the second term in the Hamiltonian keeping in mind that the theory is defined in the reduced phase space given by the constraints. In the non-abelian case it proves more useful to use $\{\,^*F_{0\alpha},\Pi^\alpha\}$ as the coordinates in phase space and look for a canonical transformation \begin{equation} \label{41bb} \{\,^*F_{0\alpha},\Pi^\alpha\}\rightarrow \{\,^*{\tilde F}_{0\alpha}, {\tilde \Pi}^\alpha\}. \end{equation} Then in order to define correctly the phase space of the theory we have to introduce first order formalism for the initial Lagrangian. The idea is to introduce a Lagrangian $L[{\tilde F},A]$, where now ${\tilde F}$ are arbitrary two-forms in the manifold, arranged to give ${\tilde F}=dA-A\wedge A$ from its equations of motion. Now the ${\tilde F}$ have no dynamical meaning since they have no time derivative, and the momenta are conjugate to the $A$-variables: \begin{equation} \label{cor1} \Pi^{am}=\frac{\delta L[{\tilde F},A]}{\delta {\dot A}^a_m}. \end{equation} It is easy to see that the following Lagrangian: \begin{equation} \label{cor2} L[{\tilde F},A]=\frac{i}{8\pi}Tr(-\frac{1}{{\bar \tau}} {\tilde F}^+_{mn}{\tilde F}^{+mn}+\frac{1}{\tau} {\tilde F}^-_{mn}{\tilde F}^{-mn}-2({\tilde F}^+_{mn}F^{+mn}- {\tilde F}^-_{mn}F^{-mn})), \end{equation} with $F=dA-A\wedge A$, is such that (\ref{4uno}) is obtained when solving the equations of motion for ${\tilde F}$. The canonical momenta are given by: \begin{eqnarray} \label{cor3} &&\Pi^{a\alpha}=\frac{\delta L[{\tilde F},A]}{\delta {\dot A}^a_\alpha} =-4\,^*{\tilde F}^{(a)0\alpha} \nonumber\\ &&\Pi^{a0}=0 \end{eqnarray} and coincide with (\ref{i1}) when substituting the equations of motion. The Hamiltonian is also given by (\ref{i2}). The canonical transformation \begin{eqnarray} \label{41bc} &&\Pi^{a\alpha}=-4\,^*{\tilde F}^{(a)0\alpha} \nonumber\\ &&{\tilde \Pi}^{a\alpha}=4\,^*F^{(a)0\alpha}, \end{eqnarray} i.e. the usual interchange between electric and magnetic degrees of freedom, produces the following ``dual'' Hamiltonian: \begin{equation} \label{corr1} {\tilde H}=\frac14 \frac{{\bar \tau}\tau}{{\bar \tau}-\tau} {\tilde \Pi}^{a\alpha}{\tilde \Pi}^a_\alpha+ \frac{{\bar \tau}+\tau}{{\bar \tau}-\tau}{\tilde \Pi}^{a\alpha} \,^*{\tilde F}^{(a)}_{0\alpha}+\frac{4}{{\bar \tau}-\tau} \,^*{\tilde F}^{(a)0\alpha}\,^*{\tilde F}^{(a)}_{0\alpha}, \end{equation} which is of the same form than the Hamiltonian of the initial theory with ${\tilde \tau}=-1/\tau$. However one must be careful with the constraints. In particular the secondary constraints (\ref{41bf}) imply for the dual theory: \begin{equation} \label{cor8} \partial_\alpha\,^*{\tilde F}^{(a)0\alpha}- f_{abc}A^b_\alpha ({\tilde F}) \,^*{\tilde F}^{(c)0\alpha}=0. \end{equation} These equations are not satisfied by the Yang-Mills theory defined from ${\tilde F}$, so although (\ref{corr1}) would naively imply that the dual theory is a Yang-Mills theory with ${\tilde \tau}=-1/\tau$, the analysis of the constraints shows that this is not the case. For the abelian theory the corresponding equation implies that ${\tilde F}$ is defined from a dual vector potential ${\tilde A}$, but the absence of a non-abelian analogue of Poincar\`e's lemma does not allow to conclude the same in the non-abelian case. Inversely if we would consider (\ref{cor8}) as an equation determining $A({\tilde F})$ we would find incompatibility with (\ref{41bc}). We can then conclude that the usual interchange between electric and magnetic degrees of freedom does not relate Yang-Mills theories with inverted couplings. Let us now obtain the ``true'' dual theory. The point is to realize that in (\ref{cor2}) we can integrate $A$ instead of ${\tilde F}$ and in this way a new theory is obtained. The equations of motion for $A$ are: \begin{equation} \label{cor5} \partial_n \,^*{\tilde F}^{(a)mn}-f_{abc} A^b_n\,^*{\tilde F}^{(c)mn} =0, \end{equation} which imply: \begin{equation} \label{ultima} A^a_m=R^{ab}_{mn}\partial_p \,^*{\tilde F}^{(b)np}, \end{equation} where $R$ is the inverse of ${\rm ad}\,^*{\tilde F}$ and it is a well defined matrix for arbitrary ${\tilde F}$ in four dimensions. Substituting in (\ref{cor2}) we get: \begin{equation} \label{cor9} {\tilde L}=\frac{i}{8\pi} (-\frac{1}{{\bar \tau}} {\tilde F}^{(a)+}_{mn}{\tilde F}^{(a)+mn}+\frac{1}{\tau} {\tilde F}^{(a)-}_{mn}{\tilde F}^{(a)-mn}+ 2 R^{ab}_{mn}(\,^*{\tilde F})\partial_q\,^*{\tilde F}^{(a)qm} \partial_p\,^*{\tilde F}^{(b)np}), \end{equation} where ${\tilde F}$ are arbitrary two-forms in the manifold\footnote{ We have not written explicitly the determinant coming from the gaussian integration.}. This Lagrangian is of Freedman-Townsend's type \cite{ft}, i.e. ${\tilde F}$ is the ``fundamental'' variable, not defined from a vector potential ${\tilde A}$, and it has been already shown to be dual to Yang Mills \cite{hfs,moha,gs}. We should remark that in order to have a well defined dual theory a prescription must be given for the pole when $R$ becomes singular \cite{gs}. The Hamiltonian associated to (\ref{cor9}) looks quite different from the Hamiltonian of Yang-Mills and it is not easy to see if they could be related by a canonical transformation. In any case such a transformation would have to map a phase space of two-forms $\{\,^*F_{0\alpha},\Pi^\alpha\}$ into a phase space of two and one forms $\{{\tilde F},{\tilde \Pi}\}$ (since the canonical momenta are conjugate to two forms), which means that if existing at all it would be quite different from an interchange between electric and magnetic degrees of freedom. Let us point out however that the first equation in (\ref{41bc}) (which maps two-forms on two-forms) produces the mapping from the constraint (\ref{41bf}) of the initial theory to the equation of motion (\ref{cor8}) of the dual. This means that after all this interchange could have some physical meaning. \vspace*{0.5cm} In this section we have been able to obtain the dual theory by manipulating the path integral in a similar way to the one known in the abelian case. One just needs to consider the intermediate Lagrangian $L[{\tilde F},A]$ and integrate out either ${\tilde F}$ or $A$ to obtain the initial or the dual Lagrangians. The group in which the dual variables live is the dual of the original group (the metric defined by its weight vectors is the inverse of the one in the original gauge group). In order to see this one needs to proceed more carefully in the previous derivation and take the original gauge fields in the fundamental representation of the gauge Lie algebra and the metric defined by the weight vectors $g_{ab}\equiv d_{ab}$. Following the steps explained above one arrives to a dual Lagrangian with metric ${\tilde g}_{ab}={\tilde d}_{ab}$ where ${\tilde d}_{ab}d^{bc}\equiv \delta^c_a$. \section{Generalization to p-forms abelian gauge theories} \setcounter{equation}{0} The generalization to $p$-forms abelian gauge theories in $d$ dimensions is direct from what we have studied in section 2\footnote{Dualization of spin-0 and spin-2 gravity-like theories has been studied in \cite{curt}.}. We are going to consider the case $d=2(p+1)$ which is the one in which both the initial and dual theories are expressed as functions of $(p+1)$-forms\footnote{In the arbitrary case the dual theory would depend on $(d-p-1)$ forms.}. The generalized S-duality transformation is implemented in the path integral by gauging the global isometry \begin{equation} \label{3uno} A\rightarrow A+\epsilon \end{equation} where now $A$ and the gauge parameter are $p$-forms. The total derivative term that gives information about the generating functional of the canonical transformation is $d{\tilde A}\wedge dA$, with ${\tilde A}$, the Lagrange multiplier, also a $p$-form. It is immediate to show that the canonical transformation is generated by the type-I generating functional\footnote{Our conventions are: $\,^*F^{i_1\dots i_{p+1}}=\frac{1}{(p+1)!}\epsilon^{i_1\dots i_d} F_{i_{p+2}\dots i_d}$ and ${\tilde F}\wedge F=\epsilon^{i_1\dots i_d} {\tilde F}_{i_1\dots i_{p+1}}F_{i_{p+2}\dots i_d}$.}: \begin{equation} \label{3dos} F=-\frac{1}{(p+1)!}\int d^dx d{\tilde A}\wedge dA \end{equation} which produces: \begin{equation} \label{3tres} \Pi^{\alpha_1\ldots\alpha_p}=\frac{\delta F}{\delta A_{\alpha_1\ldots\alpha_p}}= -((p+1)!)^2\,^*{\tilde F}^{0\alpha_1\ldots\alpha_p} \end{equation} \begin{equation} \label{3cuatro} {\tilde \Pi}^{\alpha_1\ldots\alpha_p}=-\frac{\delta F}{\delta {\tilde A}_{\alpha_1\ldots\alpha_p}}= ((p+1)!)^2\,^*F^{0\alpha_1\ldots\alpha_p} \end{equation} The same relation (\ref{217c}) for the wave functionals holds in this case since the generating functional is linear in the initial and dual variables. From it we can obtain global information about the dual variables. We can also obtain the modular weight appearing in the transformation of the partition function \cite{barbon}. Let us consider first the case $p$ odd. $p+1$ is even and then the theory allows for a $\theta$-term. In phase space (we omit the $p$ indices): \begin{equation} \label{3cinco} Z_{ps}=\int {\cal D}A {\cal D}\Pi e^{-\frac{i}{8\pi} \int d^dx({\dot A}\Pi-H)}= ({\rm Im}\tau)^{B_{p+1}/2}\int {\cal D}A e^{-S}, \end{equation} after regularizing the determinant coming from the gaussian integration on the momenta, $(p+1)$-forms in this case. The dual phase space partition function coincides with the initial one and it is given by: \begin{equation} \label{3seis} {\tilde Z}_{ps}=\int {\cal D}{\tilde \Pi}{\cal D}{\tilde A} e^{-\frac{i}{8\pi}\int d^dx ({\dot {\tilde A}}{\tilde \Pi}-{\tilde H})} =({\rm Im}\tau)^{B_{p+1}/2} \tau^{-B^-_{p+1}/2} {\bar \tau}^{-B^+_{p+1}/2} \int {\cal D}{\tilde A} e^{-{\tilde S}} \end{equation} The configuration space partition function is: \begin{equation} \label{3siete} Z=({\rm Im}\tau)^{N_p/2}\int {\cal D}A e^{-S} \end{equation} where we have followed the notation in \cite{barbon}, $N_p$ being the dimension of the space of $p$ forms after substracting all the gauge invariances (see \cite{barbon} for the detailed analysis). In the dual model the partition function is the same with $\tau\rightarrow -1/\tau$. Then we have: \begin{equation} \label{3ocho} Z=\tau^{-\frac{\chi-\sigma}{4}}{\bar \tau}^{-\frac{\chi+\sigma}{4}} {\tilde Z} \end{equation} where $\chi=2(-1)^p N_p+(-1)^{p+1}B_{p+1}$ is the Euler number and $\sigma=B^+_{p+1}-B^-_{p+1}$ the signature of the manifold. In the case $p$ even a $\theta$-term does not exist. Similar arguments to the ones above yield: \begin{equation} \label{3nueve} Z=(\frac{4\pi}{g^2})^{\chi/2}{\tilde Z} \end{equation} All these results agree with the ones presented in \cite{barbon}. \section{Conclusions} \setcounter{equation}{0} We have seen that for non-supersymmetric abelian four dimensional gauge theories S-duality can be implemented as a canonical transformation in the phase space of the theory which is the usual interchange between electric and magnetic degrees of freedom. This transformation can be generalized to the case of non-abelian gauge theories, where it seems to yield a Yang-Mills theory with inverted couplings. However in this case the canonical transformation produces some constraints in the dual theory which in the absence of a non-abelian analogue of Poincar\`e's lemma do not imply that a dual vector potential exists. The dual theory is not a Yang-Mills theory but rather a Freedman-Townsend's type of theory \cite{ft} with inverted couplings. This is shown by defining an intermediate Lagrangian depending on the initial vector potential $A$ and the dual variables ${\tilde F}$ and from which the initial and dual Lagrangians are obtained by integrating ${\tilde F}$ or $A$ respectively. It is argued that if a canonical transformation relating the two theories exists it would be far different from an interchange between electric and magnetic degrees of freedom. For the abelian case we have seen that in phase space the partition function is invariant under S-duality and it is only after integrating out the momenta degrees of freedom that a modular factor appears and the partition function in configuration space transforms as a modular function. We have generalized the canonical transformation approach to $d$-dimensional abelian gauge theories defined with $p$ forms and obtained the corresponding modular weights appearing in the transformation of the partition function. It could be very interesting to generalize the results presented in this paper to the case of supersymmetric gauge theories. \subsection*{Acknowledgements} I would like to thank O. Alvarez and N. Mohammedi for useful discussions and especially J.L.F. Barb\'on for interesting remarks leading to the final form of this paper. A Fellowship from M.E.C. (Spain) is acknowledged for partial financial support. \newpage
\section{Introduction} The Grishchuk--Zel'dovich effect gives the contribution to the microwave background anisotropy from a very large scale adiabatic density perturbation, under the standard hypothesis that this perturbation is a typical realization of a homogeneous Gaussian random field. In this paper we place upper bounds on the spectrum of density perturbations on very large scales in an open universe, corresponding to a density parameter $\Omega_0<1$ with no cosmological constant. In order to facilitate the discussion, let us make the usual assumption that the microwave background anisotropy, as well as the large scale structure, arises entirely from the spectrum of density perturbations. It is convenient to work not with the density perturbation itself, but with the primordial curvature perturbation that corresponds to it. On scales much smaller than the observable universe the spectrum of the primordial curvature perturbation can be probed directly, because there is a wealth of data relating to such scales. At the most direct level (though of course oversimplified) one can Fourier analyze the observed galaxy number density, and the primordial curvature perturbation on each scale is then related to the Fourier coefficient, up to uncertainties regarding the correct transfer function and the possibility of biased galaxy formation. In this way one finds that the primordial spectrum is approximately scale independent. On larger scales the only relevant data consist of the low multipoles of the cosmic microwave background anisotropy, and here the situation is less clear-cut. Roughly speaking these low multipoles measure the low-order spatial derivatives of the curvature perturbation, averaged over the observable universe. For this reason, the standard hypothesis is that they probe the spectrum on scales of order the size of the observable universe. If the spectrum is taken to be the scale-independent extrapolation of the one measured on smaller scales, this will indeed give a good account of the data, for either flat or open universes. The question is whether the spectrum on scales much larger than the observable universe could mimic the same effect. Could the low derivatives of the curvature perturbation, measured by the cosmic microwave background anisotropy, come from very long wavelength and large amplitude contributions to the spectrum? More generally, what {\em upper limit} can one place on the long wavelength spectrum of the curvature perturbation, by requiring that its contribution to the low multipoles be no bigger than the observed total? For a spatially flat universe this question was asked, and essentially answered, by Grishchuk and Zel'dovich \cite{GZ}. They found that a very long wavelength contribution would be present only in the quadrupole, and such a contribution has come to be known as the Grishchuk--Zel'dovich effect. On the basis of upper bounds on the quadrupole existing at that time, they showed that if the geometry distortion is of order unity on some very large scale, then that scale must be at least a factor of a hundred bigger than the observable universe. In the case of a flat universe, little has changed since their pioneering paper. Despite the fact that the large angle anisotropies have now been measured by the COBE satellite \cite{COBE}, there is no indication of a contribution from very large scales affecting the quadrupole, making it stand out relative to the higher multipoles. Consequently, one still has only a limit on how quickly the perturbation spectrum can rise on large scales and the numerical value of the scale at which the curvature perturbation can reach unity is not much changed. Now let us turn to the Grishchuk--Zel'dovich effect in the open universe, which necessarily explores scales on which spatial curvature is significant barring the exceptional case that $\Omega_0$ is very close to one. Until recently, considerable confusion existed about the correct treatment of such scales. In order to describe a cosmological perturbation one performs a mode expansion, using eigenfunctions of the comoving Laplacian so that, to first-order, each mode decouples. It has long been known to cosmologists that in the open universe any square integrable {\em function} can be generated using only {\em sub-curvature} modes, defined as those whose eigenvalue is in the range $-\infty$ to $-1$ in units of the curvature scale. Presumably with this in mind, cosmologists have retained only these modes in the expansion. But in the cosmological application we want to generate not a single function but a Gaussian {\em random field}, consisting of a set of functions together with a probability distribution, and this is done by assigning an independent Gaussian probability distribution to each coefficient in the mode expansion. (The spectrum is essentially the variance of the Gaussian distribution, and it defines the random field completely.) Cosmologists have only recently \cite{LW} discovered the fact, known to mathematicians for half a century~\cite{krein}, that the most general homogeneous Gaussian random field contains modes with eigenvalue in the entire range from $-\infty$ to $0$, including not only the sub-curvature modes but also {\em super-curvature} modes, with eigenvalues in the range $-1$ to $0$. (An illustration of the need for super-curvature modes is provided by equation (\ref{opencorr}) below; it shows that the correlation length, defined as the distance out to which the correlation function is roughly constant, can be arbitrarily large if super-curvature modes are included, whereas it is at most of order the curvature scale if these modes are omitted.) The Grishchuk--Zel'dovich effect explores by definition scales that are much bigger than the size of the observable universe, so, ignoring for the moment the exceptional case that $\Omega_0$ is very close to one, it necessarily explores very large super-curvature scales corresponding to an eigenvalue close to zero. This fact was first recognized in \cite{LW}, previous analyses \cite{turner,KS,KTF} having failed to recognize it. On the other hand, in the limit of infinitely large scales the perturbations become homogeneous and thus can be represented by superpositions of Bianchi type~V models \cite{Lukash} which are the anisotropic generalization of the open FRW metric. The microwave background anisotropy in such models has been discussed elsewhere \cite{Nature}. The present paper explores the Grishchuk--Zel'dovich effect in detail for the first time, quantifying the upper bound that can be placed on the spectrum and explaining its physical significance. \section{Mode Functions and Microwave Anisotropies} The line element for an open universe can be written as \begin{equation} ds^2 = -dt^2 + a^2(t)\left[dr^2 + \sinh^2 r \left( d\theta^2 + \sin^2 \theta \, d\phi^2 \right) \right] \,. \label{line} \end{equation} In this expression $a(t)$ is the scale factor of the universe, normalized so that the spatial curvature scalar is \begin{equation} \label{rthree} R^{(3)}=-6/a^2 \,. \end{equation} Space is practically flat on scales much less than $a(t)$ but strongly curved on much larger scales, so one may call the distance $a(t)$ the physical {\em curvature scale}. In these units, the comoving curvature scale is unity. The Friedmann equation can be written \begin{equation} \label{omega} 1-\Omega= \frac 1{(aH)^2} \,, \end{equation} where as usual $H=\dot a/a$ is the Hubble parameter, and the time-dependent quantity $\Omega$ is the energy density measured in units of the critical density $3H^2/8\pi G$. From this equation it is clear that the Hubble distance $H^{-1}$ is always less than the curvature scale. It is useful to employ conformal time, defined by $\eta \equiv \int dt/a(t)$, which has the interpretation of being the coordinate distance to the particle horizon. Its present value $\eta_0$ is an excellent approximation to the distance to the last scattering surface and, making the approximation (valid except for very low $\Omega_0$) that the universe can be treated as matter dominated since last scattering, it is given by \begin{equation} \eta_0 = \cosh^{-1} \left[ \frac{2-\Omega_0}{\Omega_0} \right] \,, \end{equation} where subscript `0' always indicates present value. For $\Omega_0 < 2/(1 + \cosh 1) \simeq 0.786$, the surface of last scattering is located beyond the curvature scale. In analyzing the Grishchuk--Zel'dovich effect, one needs the full paraphernalia of mode functions appropriate to an open universe. This has recently been given in considerable detail by Lyth and Woszczyna \cite{LW}, and we shall simply repeat the formulae here. The mode expansion of a generic function $f$ is \begin{equation} f(r,\theta,\phi,t)=\int_0^\infty \mbox d k \sum_{lm} f_{klm}(t) \Pi_{kl}(r) Y_{lm}(\theta,\phi) \,, \end{equation} Here $-(k/a)^2$ is the eigenvalue of the Laplacian corresponding to the eigenfunction $\Pi_{kl}(r) Y_{lm}(\theta,\phi)$. For brevity we shall refer to $k/a$ as the wavenumber, even though one cannot usefully define plane waves in curved space. We introduce the notation $q^2=k^2-1$; sub-curvature modes correspond to $0 < q^2 < \infty$ and super-curvature modes correspond to $-1 < q^2 < 0$. The angular functions are the usual spherical harmonics. The radial functions $\Pi_{kl}$ for the sub-curvature modes are given by \cite{HAR} \begin{eqnarray} \Pi_{kl} & \equiv & N_{kl} \tilde \Pi_{kl} \,, \\ \tilde \Pi_{kl} & \equiv & q^{-2} (\sinh r)^l\left(\frac{-1}{\sinh r} \frac{{\mbox d}}{{\mbox d} r}\right)^{l+1} \cos(q r) \,, \\ N_{kl} & \equiv & \sqrt\frac2\pi q^2 \prod_{n=0}^l (n^2+q^2)^{-1/2} \,. \end{eqnarray} For the super-curvature modes they can be obtained by analytic continuation \cite{LW} \begin{eqnarray} \Pi_{kl} & \equiv & N_{kl}\tilde\Pi_{kl} \,, \\ \tilde \Pi_{kl} & \equiv & |q|^{-2} (\sinh r)^l\left(\frac{-1}{\sinh r} \frac{{\mbox d}}{{\mbox d} r}\right)^{l+1} \cosh(|q| r) \,, \\ N_{kl} & \equiv & \sqrt\frac2\pi |q| \prod_{n=1}^l (n^2+q^2)^{-1/2} \hspace{3em} (l\geq1) \,. \end{eqnarray} To construct a Gaussian random field the amplitude $f_{klm}$ of each mode is given an independent Gaussian probability distribution, whose variance is defined by the spectrum ${\cal P}_f(k)$ through \begin{equation} \label{fullspec} \langle f^*_{klm} f_{k'l'm'}\rangle = \frac{2\pi^2}{k|q|^2} {\cal P}_f(k) \delta(k-k') \delta_{ll'} \delta_{mm'} \,, \end{equation} where the angle brackets denote the ensemble average. It determines the two-point correlation function through the relation \begin{equation} \label{opencorr} \langle f(1)f(2) \rangle = \int^\infty_0 \frac{\mbox d k}{k} {\cal P}_f(k) \frac{\sin(q r)}{q\sinh r} \,. \end{equation} The correlation function depends only on the geodesic distance $ar$ between the comoving points 1 and 2 provided that the spectrum is independent of $l$ and $m$, and the Gaussian field is then said to be homogeneous (with respect to the group of transformations leaving the distance invariant). Evaluated at $r=0$ it gives the position-independent mean-square \begin{equation} \label{mssc} \langle f^2 \rangle = \int_0^\infty \frac{{\mbox d} k}{k}{\cal P}_f(k) \,. \end{equation} We are interested in the perturbation $\delta R^{(3)}_{klm}$ in the curvature scalar of comoving hypersurfaces (those orthogonal to comoving worldlines), away from its average value $6/a^2$. It is conveniently characterized by a quantity ${\cal R}$ defined by \begin{equation} 4(k^2+3) {\cal R}_{klm}/a^2 =\delta R^{(3)}_{klm} \,. \label{rdef} \end{equation} After matter domination (which is the only era that concerns us) ${\cal R}$ is practically constant until $\Omega$ breaks away from 1. Before that happens, ${\cal R}$ is equal to $- 5/2$ times the gauge invariant gravitational potential $\Phi$ \cite{Bardeen}, but afterwards $\Phi$ is multiplied by a factor $F(\eta)$ defined by \begin{equation} \label{FETA} F(\eta) = 5\,\frac{\sinh^2\eta-3\eta\sinh\eta+4\cosh\eta-4} {(\cosh\eta-1)^3}\,. \end{equation} As long as it is constant, ${\cal R}$ is related to the density perturbation on comoving hypersurfaces by \begin{equation} \frac{\delta\rho}{\rho}= \frac{2}{5} \, \frac{k^2+3}{a^2H^2} \, {\cal R}_{klm} \,. \label{denspert} \end{equation} {}From now on $\cal R$ will refer exclusively to the constant, early time value. The only important effect on the cosmic microwave background anisotropy from large scales will be gravitational, through the usual Sachs--Wolfe effect. As the usual practice is to work with the monopole and dipole subtracted from measured anisotropies, we shall concentrate on the multipoles from the quadrupole ($l=2$) upwards. However, one must also consider the question of whether or not the dipole induced by the perturbations is compatible with observations, and we discuss this in the Appendix. Following Lyth and Woszczyna \cite{LW}, the appropriate expression (for $l \geq 2$) is \begin{equation} \label{openswsc} l(l+1)C_l = 2\pi^2 l(l+1) \int^\infty_0 \frac{{\mbox d} k}{k} {\cal P}_{\cal R}(k) I_{kl}^2 \ , \end{equation} where $C_l$ is the radiation angular power spectrum (defined as usual as the ensemble average of the $l$-th multipole of the temperature anisotropy), and ${\cal P}_{\cal R}$ is the spectrum of the time-independent primordial curvature perturbation. The function $I_{kl}^2$ is the `window function' which indicates how a given scale $k$ contributes to the $C_l$. It is given by \begin{equation} \label{window} q I_{kl} = \frac{1}{5} \,\Pi_{kl}(\eta_0) + \frac{6}{5} \int^{\eta_0}_0 \mbox d r \,\Pi_{kl}(r)\,F'(\eta_0-r) \,, \end{equation} where in calculating the $I_{kl}$, one must use the mode expansion appropriate to whether the mode is sub-curvature or super-curvature. In order to uncover the complete form of the window functions for $\Omega_0 < 1$ it is necessary to evaluate them numerically, both for sub- and super-curvature modes. We show the window functions corresponding to the first three multipoles in Figure 1, for the case $\Omega_0 = 0.2$. As expected, they behave smoothly across the curvature scale $k = 1$. The window functions show that if the spectrum continues to be flat or falling as one goes to larger scales, then there will be little effect from the large scale modes, and so the large angle cosmic microwave anisotropy will indeed be dominated by modes of order the Hubble scale. Only if the spectrum rises can the effect of very large scales be significant. This is true regardless of the value of $\Omega_0$. \section{The sub-curvature scale Grishchuk--Zel'dovich effect} We are interested in the effect on the microwave anisotropies of scales far bigger than the observable universe. If $\Omega_0$ is close to one the curvature scale becomes large compared with the size of the observable universe, giving the possibility that the large scales can still be in the sub-curvature regime. We consider this first, moving on to the question of the super-curvature regime in the following Section. We shall see that in the limit $\Omega_0 \to 1$, only the sub-curvature effect is important. As we have chosen the scale factor $a$ to be equal to the curvature scale, the limit $\Omega_0=1$ corresponds to the limit $a \to \infty$. Since a physical wavenumber is $k/a$ and a physical radial distance is $ar$, it also corresponds to $k \to \infty$ and $r \to 0$ with $kr$ fixed. One can identify $q$ with $k$ in this limit, and the radial functions are related to the spherical Bessel functions by \begin{equation} \label{jlim} \Pi_{kl}( r)\to \sqrt\frac{2}{\pi} \, k j_l(k r) \,. \end{equation} The correlation function expression becomes the usual \begin{equation} \label{flatcorr} \langle f(1)f(2) \rangle = \int^\infty_0 \frac{\mbox d k}{k} {\cal P}_f(k) \frac{\sin(k r)}{kr} \,, \end{equation} and the relation between ${\cal R}$ and the curvature scalar becomes \begin{equation} \label{tweiflat} 4 k^2 {\cal R}_{klm}/a^2 = \delta R^{(3)}_{klm} \,. \end{equation} The integral in Eq.~(\ref{window}) vanishes and one finds \begin{equation} \label{clflat} l(l+1) C_l= \frac{4\pi}{25} l(l+1) \int_0^{\infty} \frac{\mbox d k }{k} \, {\cal P}_{\cal R}(k)\, j_l^2 (\eta_0 k) \,. \end{equation} If the curvature perturbation spectrum is scale independent this gives a constant value of $l(l+1)C_l$ \begin{equation} l(l+1)C_l=\frac{2\pi}{25}{\cal P}_{\cal R} \,, \end{equation} which is consistent with the COBE data \cite{COBE}. We represent the effect of large scale modes by a delta function contribution to the spectrum. We consider a very large scale {\em sub-curvature} mode \begin{equation} \label{vvlarge} {\cal P}_{{\cal R}}^{\rm SUB} \simeq \delta(\ln k- \ln k_{\rm SUB}) \langle {\cal R}^2 \rangle _{\rm SUB} \,, \end{equation} where \begin{equation} 1 \ll k_{\rm SUB} \ll a_0H_0= (1-\Omega_0)^{-1/2} \,. \label{ksub} \end{equation} Since $j_l(x) \propto x^l$ as $x \to 0$, the quadrupole dominates all other multipoles and is given by \begin{equation} \label{subquad} 6 C^{\rm SUB}_2 = \frac{148\pi}{1875} \left(\frac{k_{\rm SUB}}{a_0H_0}\right)^4 \langle {\cal R}^2 \rangle_{\rm SUB} \,. \end{equation} This, in more modern and precise notation, is the result derived by Grishchuk and Zel'dovich \cite{GZ} for the case $\Omega_0=1$. In the COBE data the quadrupole is not markedly higher than the others (quite the reverse if anything), so we take the observational value of $l(l+1)C_l$ as an upper limit on the effect. This gives \begin{equation} \label{climit2} l(l+1) C_l^{\rm SUB}<8 \times 10^{-10} \,. \end{equation} We have taken the observational limit as that corresponding to a flat spectrum evaluated for $\Omega_0=1$ with an upper limit of $20\mu$K on the expected quadrupole $Q = \sqrt{5C_2/4\pi}$ \cite{Gorski}. This leads to the bound \begin{equation} \label{omega1con} \left( \frac{k_{{\rm SUB}}}{a_0H_0}\right)^2 \langle {\cal R}^2 \rangle^{1/2}_{{\rm SUB}} < 6 \times 10^{-5}\,. \end{equation} This bound was derived under the assumption that $k_{{\rm SUB}}$ is a large {\em sub-curvature} scale, $k_{{\rm SUB}} \gg 1$, which is consistent with Eq.~(\ref{omega1con}) only if $1-\Omega_0 \lesssim 10^{-4} \langle {\cal R}^2 \rangle^{-1/2}_{{\rm SUB}}$. As we shall see in Section \ref{phys}, the biggest permissible geometry distortion corresponds to $\langle {\cal R}^2 \rangle_{{\rm SUB}}\sim 1$, and with this value consistency requires $1-\Omega_0 \lesssim 10^{-4}$. In words, a curvature perturbation of order unity on sub-curvature scales is allowable only if $\Omega_0$ is very close to one \cite{KTF}. This is to be expected, since otherwise there are no sub-curvature scales much bigger than the observable universe. \section{The super-curvature scale Grishchuk--Zel'dovich effect} If $\Omega_0$ is not close to one, scales far larger than the observable universe are necessarily much bigger than the curvature scale, corresponding to super-curvature modes with $0<k\ll 1$. We again consider a delta function power spectrum, given by \begin{equation} \label{dfps} {\cal P}_{\cal R}^{\rm VL} \simeq \delta(\ln k-\ln k_{\rm VL}) \langle {\cal R}^2 \rangle_{\rm VL} \,, \end{equation} where $k_{\rm VL}$ is a scale satisfying $0<k_{\rm VL} \ll 1$. The effect of such a contribution was investigated qualitatively in Refs.~\cite{LW,Munich}, but here we present a full quantitative calculation and also explain more fully the physical significance of the result. The limit of small $k$ can be taken partly analytically, though one can obtain the same results by numerical calculation from the full expressions given above. In this limit, $\Pi_{k0} \to 1$, but the other radial functions are proportional to $k$. It is convenient to define \begin{eqnarray} N_l &\equiv& \lim_{k\to0}\ k\,N_{kl} \ ,\\ \tilde\Pi_l &\equiv& \lim_{k\to0}\ \tilde\Pi_{kl}/k^2 \ , \end{eqnarray} from which one finds \begin{equation} N_l = \sqrt{\frac2\pi} \,\prod_{n=2}^l(n^2-1)^{-1/2} \hspace{5mm}(l\geq 2) \ . \end{equation} and \begin{eqnarray} \label{PI12} \tilde\Pi_{1}(r) &=& {1\over2}\left[\coth\,r - {r\over\sinh^2 r}\right] \,, \\ \tilde\Pi_{2}(r) &=& {1\over2}\left[1 + {3(1 - r\,\coth\,r)\over\sinh^2 r}\right] \,. \end{eqnarray} The other radial functions follow from the recurrence relation \begin{equation} \tilde\Pi_{l}(r) = - l(l-2)\,\tilde\Pi_{l-2}(r) + (2l-1)\,\coth\,r\ \tilde\Pi_{l-1}(r)\,. \end{equation} Using these results, the contribution to the mean square multipoles becomes \begin{equation} \label{clvl} l(l+1)\,C_l^{\rm VL} = l(l+1)\,N_l^2\,B_l^2\, k^2_{\rm VL} \langle{\cal R}^2\rangle_{\rm VL} \,, \end{equation} where \begin{equation} B_{l} \equiv \frac{1}{5} \, \tilde\Pi_l(\eta_0)+\frac{6}{5} \, \int^{\eta_0}_0 {\rm d}r\,\tilde\Pi_l(r)\,F'(\eta_0- r)\ . \end{equation} By evaluating the full expressions in Eq.~(\ref{openswsc}) numerically, we have found that these limits are an excellent approximation for calculating $C_l$, at least for the low multipoles we consider. For $k_{\rm VL} \leq 0.1$ the approximation is good to within a few percent for any reasonable $\Omega_0$. \subsection{The shape of the spectrum} In the case $\Omega_0=1$ the Grishchuk--Zel'dovich effect comes entirely from the sub-curvature modes that we discussed in the previous Section, and is present only in the quadrupole. It was noted in Refs.~\cite{LW,Munich} that for $\Omega_0<1$, on the other hand, the super-curvature Grishchuk--Zel'dovich effect is present in all multipoles\footnote{As $\Omega_0$ becomes small, the actual pattern produced by a single infinite scale mode becomes a hot (or cold) spot on the sky of angular size $\sim 80\, \Omega_0$ degrees \cite{Nature}.}. However, in those papers the $l$ dependence was not evaluated. For $\Omega_0$ close to one, it can be shown analytically that $C_l^{\rm VL} \propto (1-\Omega_0)^l$, so that the quadrupole dominates.\footnote{This result that the quadrupole dominates refers to the very large scale super-curvature modes corresponding to $k\ll 1$. In the previous Section we found that the quadrupole also dominates for large scale sub-curvature modes corresponding to $1\ll k\ll a_0H_0$. It also dominates for the intermediate scales $k \sim 1$; the origin of this result is that for small $r$, both sub- and super-curvature modes have the same leading behaviour when expanded in small $|q|r$.} The prefactor can be evaluated analytically, yielding \begin{eqnarray} \label{C2KR} 6 C_2^{\rm VL} & \simeq & \frac{64}{625\pi} \, k^2_{{\rm VL}} \langle{\cal R}^2\rangle_{\rm VL} \, (1-\Omega_0)^2 \,, \\ \label{C3KR} 12 C_3^{\rm VL} & \simeq & \frac{4096}{30625\pi} \, k^2_{{\rm VL}} \langle{\cal R}^2\rangle_{\rm VL} \, (1-\Omega_0)^3 \,. \end{eqnarray} We have confirmed that our numerical code reproduces these coefficients. It demonstrates that these expressions are very accurate for $\Omega_0 > 0.9$, and still a reasonable approximation for $\Omega_0 = 0.8\,$. For $\Omega_0$ significantly below one, numerical calculation is essential. In Figure 2 we show the shape of the radiation power spectra for several different values of $\Omega_0$, normalized such that all have the same $C_2$. One sees a complicated behaviour, which is not so surprising since standard calculations of the multipoles using spectra of sub-curvature modes already show a very complex behaviour with $\Omega_0$ even if only the Sachs-Wolfe terms are included \cite{KS,KRSS}. Unusual behaviour is particularly marked for $\Omega_0 \simeq 0.35$; an `accidental' suppression of the quadrupole even from curvature-scale perturbations, due to cancellation between the `intrinsic' and `line-of-sight' terms, has already been noted \cite{KTF}. For this $\Omega_0$ the contribution of very large scales to the octopole is considerably larger than that to the quadrupole. A question one would like to address is whether for any $\Omega_0$ the shape induced by the super-curvature modes resembles the shape observed by COBE. A convenient way of characterizing the observed shape is to note that it is compatible with the one induced in an $\Omega_0=1$ universe by power-law spectra of density perturbations, with the spectral index $n$ roughly constrained to a range 0.7 to 1.4 about the Harrison--Zel'dovich value $n=1$ \cite{Gorski}. As we have remarked, the assumption of $n=1$ leads to $l(l+1) C_l$ being constant. We have illustrated the allowed slopes of the spectra in Figure 2 for comparison. We find that over very short ranges of multipoles it is possible for the Grishchuk--Zel'dovich effect to mimic the observations, which is not possible in a flat universe. However, for no value of $\Omega_0$ is the spectrum sustained as approximately flat even over just the limited range up to $l = 10$. We conclude then that the Grishchuk--Zel'dovich effect cannot be responsible for the entire shape of the radiation power-spectrum (a conclusion which will surprise nobody!), but that it remains possible for some fraction of the measured multipoles to be due to this effect. In the flat space limit only the quadrupole can be partly generated this way. \subsection{The amplitude of the spectrum} Earlier papers \cite{LW,Munich} assumed that the quantities $N_l$ and $B_l$ are roughly of order one for low multipoles. However, we find considerable cancellations in $B_l$ and in fact their typical sizes are somewhat smaller. Continuing with the choice of a delta function power spectrum, we evaluate them numerically and compare the anisotropies to the observational limit, demanding again that they be less than the observed value \begin{equation} l(l+1) C_l^{\rm VL} <8 \times 10^{-10} \,. \label{climit} \end{equation} In Figure 3 we plot the dependence of the $l = 2$, $3$, $4$ and $5$ multipoles on $\Omega_0$. In accordance with Eq.~(\ref{clvl}), $l(l+1)C_l$ scales with $k^2_{\rm VL} \langle{\cal R}^2\rangle_{\rm VL}$, so we have plotted it normalized by this quantity. Near $\Omega_0 = 1$, the analytic approximations of Eqs.~(\ref{C2KR}) and (\ref{C3KR}) can be used. By taking the upper limits on the Grishchuk--Zel'dovich effect given above, one can obtain a maximum permitted value for $k^2_{\rm VL} \langle{\cal R}^2\rangle_{{\rm VL}}$, as a function of $\Omega_0$, and this is plotted in Figure 4. We have imposed the constraint on multipoles up to $l = 6$ rather than just the quadrupole, since the `accidental' cancellation of the quadrupole for $\Omega_0 \simeq 0.35$ would otherwise distort the constraint. We see from Figure 4 that for $0.25 \lesssim \Omega_0 \lesssim 0.8$ the bound is fairly constant \begin{equation} k^2_{\rm VL}\langle{\cal R}^2\rangle_{\rm VL} \lesssim 10^{-6} \,. \label{omegasmall} \end{equation} For $\Omega_0<0.25$ it becomes more severe. On the other hand, for $0.8 \lesssim \Omega_0 < 1$, the dependence of Eq.~(\ref{C2KR}) yields \begin{equation} k_{{\rm VL}}^2 \langle{\cal R}^2\rangle_{{\rm VL}} < \frac{3 \times 10^{-8}} {(1-\Omega_0)^2} \,. \label{omegabig} \end{equation} These bounds were derived under the assumption that $k_{{\rm VL}} \ll 1$, so Eq.~(\ref{omegasmall}) is automatically satisfied if $\langle{\cal R}^2\rangle_{{\rm VL}} \lesssim 10^{-6}$ and Eq.~(\ref{omegabig}) is automatically satisfied if $\langle {\cal R}^2 \rangle_{{\rm VL}} \lesssim 10^{-8}/(1-\Omega_0)^2$. As we shall see later, the biggest permissible geometry distortion corresponds to $\langle{\cal R}^2\rangle_{{\rm VL}} \sim 1$. Even with this maximal value, Eq.~(\ref{omegabig}) is automatically satisfied if $1-\Omega_0 \lesssim 10^{-4}$; in this regime, the maximal geometry distortion is allowed on all super-curvature scales with $k_{{\rm VL}} \ll 1$. This result is hardly surprising, since super-curvature scales move off to infinity in the limit $\Omega_0 \to 1$. The bound can also be written in terms of $k/(aH)$, which is the physical wavenumber in Hubble units. For $0.25 \lesssim \Omega_0 \lesssim 0.8$ it becomes \begin{equation} \label{omegasmall2} \frac{k^2_{\rm VL}}{a_0^2 H_0^2} \langle{\cal R}^2\rangle_{\rm VL} \lesssim 10^{-6} (1-\Omega_0) \,, \end{equation} and for $0.8 \lesssim \Omega_0 < 1$ it becomes \begin{equation} \label{omegabig2} \frac{k^2_{\rm VL}}{a_0^2 H_0^2} \langle{\cal R}^2\rangle_{{\rm VL}} \lesssim \frac{3 \times 10^{-8}}{1-\Omega_0} \,. \end{equation} Provided $\Omega_0$ is not too close to one, then, assuming the maximal value $\langle{\cal R}^2\rangle_{{\rm VL}} \sim 1$, these results say that that the inverse wavenumber must be at least about a thousand times the Hubble distance (for $\Omega_0 > 0.25$). As $\Omega_0$ approaches one, the bound weakens slowly. At $1-\Omega_0 \simeq 10^{-4}$, it requires the inverse wavenumber to be at least a factor of about a hundred times the Hubble distance, and by this time super-curvature scales are far enough outside the observable universe to automatically satisfy this requirement. \section{Physical interpretation of the bounds} \label{phys} In order to interpret these bounds physically, we have to explain the meaning of the curvature perturbation ${\cal R}$, showing that within the context of a homogeneous Gaussian random field it can be at most of order one. For this physical interpretation, we assume that the observed curvature perturbation is a typical realization of a homogeneous Gaussian random field not only within the observable universe, but also in a far bigger region, much bigger in fact than the correlation length. Even though the background space has negative curvature, the presence of a large perturbation can render regions of the universe positively curved. Moreover, within a region no bigger than the correlation length the curvature is practically homogeneous so that we are dealing with a practically homogeneous space just as in the unperturbed universe. But in a homogeneous space with positive curvature scalar $R^{(3)}$, the biggest sphere that can be drawn has radius $d$ given by\footnote{This equation follows from the positive-curvature version of Eq.~(\ref{line}), given by $\sinh\to \sin$, corresponding to the maximum value $r=\pi$. When it is small, $R^{(3)} d^2$ is a linear measure of the geometry distortion within the sphere so that for instance the area of the sphere is equal to $(1-2R^{(3)} d^2)$ times its Euclidean value $4\pi d^2$.} \begin{equation} R^{(3)} d^2=\pi^2/6\sim 1 \,. \end{equation} In other words, positive curvature inside a given sphere cannot exceed this value. One can say that for bigger positive curvature, space would `close in on itself', indicating that our underlying assumptions break down. What we are going to do is demonstrate that, at least for a perturbation coming solely from a single scale, this requirement is equivalent to ${\cal R}\lesssim 1$. The situation is different for sub- and super-curvature scales, and we begin by considering the former. In that case the correlation length $d_{\rm VL}$, defined as the distance $ar$ within which correlation function Eq.~(\ref{flatcorr}) is practically constant, is simply the inverse wavenumber \begin{equation} d_{\rm SUB}=a\, k_{\rm SUB}^{-1} \,. \end{equation} Within a typical sphere of this radius ${\cal R}$ is practically constant, and the curvature scalar given by Eq.~(\ref{rdef}) is \begin{equation} 4 {\cal R}= d_{\rm SUB}^2 \delta R^{(3)}\, . \end{equation} If ${\cal R}\sim 1$ the perturbation $\delta R^{(3)}$ always dominates the background value $R^{(3)}=-6/a^2$ because $k_{\rm SUB}\gg 1$. According to the linear cosmological perturbation theory that we are invoking in this paper, the actual value of $\cal R$ within a randomly located sphere of fixed radius has a Gaussian probability distribution.\footnote{In referring to a `randomly located' sphere we are taking for granted the ergodic property, which is essentially the statement that choosing a randomly located sphere for a fixed realization of the random field is equivalent to choosing a random realization of the field at a fixed location. Only in the latter case is the Gaussian property an immediate consequence of our assumptions. However the ergodic property can be proved \cite{adler} under weak assumptions for a Euclidean geometry corresponding to $\Omega_0=1$, and we see no reason to suppose that it fails for the case of curved space.} Taken literally the linear theory therefore predicts the existence of regions in which ${\cal R}\gg 1$, in contradiction with the physical interpretation of that quantity. It follows that at least within the framework that we are adopting the mean square contribution $\langle {\cal R} \rangle^2_{\rm SUB}$ cannot exceed one.\footnote{In what follows we assume that linear cosmological perturbation theory is valid right up to this maximum value, or in other words that it is valid as long as it predicts only rare regions of space with the unphysical value ${\cal R} \gtrsim 1$. This seems very likely because there is no obvious criterion which would indicate earlier non-linearity. In particular, on the large scales that we are considering the fractional density perturbation Eq.~(\ref{denspert}) on comoving hypersurfaces is small if ${\cal R}$ is small.} With this maximal value, Eq.~(\ref{omega1con}) becomes \begin{equation} d_{\rm SUB} \gtrsim 130 \, H_0^{-1} \,. \end{equation} This discussion of the sub-curvature case has led to no surprises. Consider now the super-curvature case, corresponding to the scale $k_{\rm VL}\ll 1$. {}From Eq.~(\ref{opencorr}) it follows that the correlation length is now given by \begin{equation} d_{\rm VL}=a_0 \, k_{\rm VL}^{-2} \,. \end{equation} This differs from the Euclidean result through the appearance of $k^2$ instead of $k$. As pointed out in \cite{LW}, the different power can be understood from the different relation between volume and area.\footnote{In Eq.~(50) of Ref.~\cite{LW}, the symbol $f$ should be replaced by $\xi$.} Within a sphere whose radius is of order $d_{\rm VL}$, the correlation length ${\cal R}$ is practically constant and so is the corresponding curvature scalar given by Eq.~(\ref{rdef}) as \begin{equation} 12 {\cal R} \simeq a^2 \delta R^{(3)} \,. \end{equation} Since the background curvature is $R^{(3)}=-6/a^2$ we see that $2{\cal R}$ is equal to the {\em fractional} change in the geometry distortion. It will be negative in some regions and positive in others, but on scales much bigger than the curvature scale the background distortion is huge (the ratio of area to volume is exponentially large compared with the Euclidean ratio). As a result the maximum allowed positive value of ${\cal R}$ is again of order one and we again require for the mean square $\langle {\cal R}^2 \rangle_{{\rm VL}} \lesssim 1$. But this maximal value now corresponds to a huge distortion of the geometry. For the maximal super-curvature geometry distortion $\langle{\cal R} \rangle^2_{\rm VL} \sim 1$, the correlation length must satisfy \begin{equation} d_{\rm VL}\gtrsim 10^6 \, H_0^{-1} (1-\Omega_0)^{-1/2} \,, \end{equation} if $0.25 \lesssim \Omega_0 \lesssim 0.8$, and \begin{equation} \label{supcons} d_{\rm VL} \gtrsim 4\times 10^7 \, H_0^{-1}(1-\Omega_0)^{3/2} \,, \end{equation} if $\Omega_0 \gtrsim 0.8$. As stated earlier, this second bound is automatically satisfied in the extreme case $1-\Omega_0\lesssim 10^{-4}$. To summarize, if there is a maximal geometry distortion on a super-curvature scale, then this scale must be more than a million times the Hubble distance if $\Omega_0=0.25$. As $\Omega_0$ rises the constraint weakens, but only in the extreme regime $1-\Omega_0\lesssim 10^{-4}$ does it disappear completely, so that all super-curvature correlation lengths $d_{\rm VL}$ are allowed. This is just the regime in which one can have maximal geometry distortion also on sub-curvature scales, provided that the correlation length is more than a hundred times the Hubble distance. \section{Discussion} While it remains possible that some component of the observed cosmic microwave background anisotropies is due to the Grishchuk--Zel'dovich effect, there is no evidence that this is the case and so at present one is left with a null result. In absolute generality, this null result tells us nothing at all about the nature of the universe beyond our horizon. This is because the observed anisotropies are due entirely to the spatial and temporal variation of the curvature perturbation at or within the last scattering surface; anything could happen to it immediately outside our observable universe without any effect being noticeable. In order to say anything further, one must adopt the hypothesis that even far beyond the observable universe the curvature perturbation is a typical realization of a homogeneous Gaussian random field. In that case one expects to be able to break the perturbation up into modes {\em including} very large scale modes, and the Grishchuk--Zel'dovich effect then dictates how rapidly the perturbations may reach large amplitudes as one goes to large scales. Although there is no way of testing whether the random field hypothesis is actually correct, the idea that the curvature perturbation is of such a form is the basis for most work in large scale structure. It makes particular sense in the case of a flat universe; remembering that the horizon scale at last scattering is much smaller than at present, it would be strange for the hypothesis to only break down at precisely the present epoch. In the case of an open universe, the hypothesis seems {\em a priori} more restrictive, because the geometry has selected out a special scale, the curvature scale, beyond which different physics might apply. While our results have been framed in a particularly general manner, it is interesting to consider their implications in the context of open universe inflationary scenarios. The motivation in the early papers on the open universe Grishchuk--Zel'dovich effect \cite{turner,KS,KTF} was an attempt to render unlikely the prospect of open universe inflation, which at that time was modeled as chaotic inflation of sufficiently short duration that the universe was not forced to spatial flatness \cite{ELM}. This was on the grounds that such inflation could not explain the homogeneity of the universe in a comoving region bigger than the Hubble distance at the beginning of inflation (and hence on any comoving scale where the curvature is significant). However, in the strictest sense as discussed above, the absence of the Grishchuk--Zel'dovich effect cannot be construed as evidence that the universe is indeed homogeneous in such a region. These papers also failed to provide a complete treatment; Turner \cite{turner} ignored spatial curvature and none of Refs.~\cite{turner,KS,KTF} included super-curvature modes. More interesting are the recently discussed `single-bubble' models of open inflation \cite{hist,STYY,STY,BGT,LM,YST}, which are capable of erasing large scale perturbations even above the curvature scale while still resulting in an open universe. This undermines the original motivation of the early work on the Grishchuk--Zel'dovich effect, but at the same time offers a new motivation, because it appears possible for super-curvature modes to be generated in such models. Indeed, in Ref.~\cite{STY} a primordial spectrum for such models is calculated and, provided the inflaton mass is sufficiently small, it includes a single discrete super-curvature mode (which approaches $k = 0$ in the massless limit) as well as a continuum of sub-curvature modes. This ties in nicely with our delta-function analysis. In Ref.~\cite{YST} the microwave anisotropies for such a composite spectrum are calculated, and as far as we can ascertain from their figures their results are in good agreement with ours. In conclusion then, we have confirmed that, even in an open universe, the low multipoles of the observed microwave anisotropies cannot be due solely to an adiabatic density perturbation on scales much larger than the observable universe, given the usual assumption that the perturbation is a typical realization of a homogeneous Gaussian random field. We have also provided limits as to how quickly the typical amplitude of perturbations can reach unity as one moves to larger scales. Although the entire observed anisotropies cannot be due to the Grishchuk--Zel'dovich effect, it remains possible that some component of them does have such an origin. This may be the case in the `single-bubble' models of open inflation. \section*{Acknowledgements} JGB and DW are supported by PPARC (UK) and ARL by the Royal Society. We thank John Barrow and Leonid Grishchuk for discussions.
\section*{Figure Captions} \noindent Figure 1: Ratio of non-equilibrium density to equilibrium density and temperature as functions of time at SPS, RHIC and LHC energies. \end{document} 
\subsection{Dispersion relation for the $NN\to NN$ amplitude} For $NN\to NN$ and $N\overline N \to N\overline N$ scattering the field theoretical scattering amplitude $T$ is related to the standard $S$-matrix by \begin{equation} S_{fi}=\delta_{fi}-i(2\pi)^{-2}\delta^{(4)}(p_1'+p_2'-p_1-p_2) \left ( \frac{m_N^4}{E_1'E_2'E_1E_2} \right) ^{\frac{1}{2}} T_{fi} \ . \end{equation} If we neglect isospin for the moment, the $s$-channel ($NN\to NN$) amplitude can be written as \begin{equation} T_s(p_1',p_2';p_1,p_2)=\overline u(p_1',\lambda_1') \overline u(p_2',\lambda_2') \hat T u(p_1,\lambda_1)u(p_2,\lambda_2) \quad, \label{invampI} \end{equation} \noindent where \begin{equation} u(p,\lambda)=\sqrt{{E(p)+m_N\over 2m_N}}\left ( \begin{array}{c} 1 \\{2\lambda p \over E(p)+m_N}\end{array}\right ) |\lambda> \end{equation} is a Dirac helicity spinor normalized to $\overline u u =1$. (The spin dependence is suppressed on the left hand side of Eq.~(\ref{invampI}).) For on-shell scattering, $\hat T$ can be expressed as a linear combination of five invariant operators $\hat C_j$; the expansion coefficients $c_j$ are scalar functions of the Mandelstam variables $s\equiv (p_1+p_2)^2$ and $t\equiv (p_1'-p_1)^2$. ($u$ is not independent, but given by $u=4m_N^2-s-t$.) $\hat T$ can then be written as \begin{equation} \hat T=\sum^5_{j=1} c_j(t,s)\hat C_j \ . \label{expans} \end{equation} In contrast to our former work dealing with correlated $\pi\pi$ exchange \cite{Kim94} we now use, instead of the so-called perturbative invariants (see Ref.~\cite{Kim94}) the Fermi-invariants \begin{eqnarray} S &=& (I)^{(1)}\, (I)^{(2)} \nonumber \\ P &=& (\gamma_5)^{(1)}\, (\gamma_5)^{(2)} \nonumber \\ V &=& (\gamma^\mu)^{(1)}\, (\gamma_\mu)^{(2)} \nonumber \\ A &=& (\gamma_5\gamma^\mu)^{(1)}\, (\gamma_5\gamma_\mu)^{(2)} \nonumber \\ T &=& (\sigma^{\mu\nu})^{(1)}\, (\sigma_{\mu\nu})^{(2)} \ \ . \label{Fermi} \end{eqnarray} \noindent where $\sigma_{\mu\nu}\equiv\frac{i}{2}[\gamma_\mu,\gamma_\nu]$. Correspondingly, the amplitude for the $t$-channel ($N\overline N \to N \overline N$) process defined in Fig.~\ref{figone} then reads \begin{equation} T_t(-p_1',p_2';p_1,-p_2)=\overline v(-p_1',\bar\lambda_1') \overline u(p_2',\lambda_2') \hat T v(p_1,\lambda_1)u(-p_2,\bar \lambda_2) \ . \label{invampII} \end{equation} Here \begin{equation} v(- p,\lambda)=\sqrt{{E(p)+m_N\over 2m_N}} \left ( \begin{array}{c} {p\over E(p)+m_N} \\ -2\lambda\end{array}\right ) |-\lambda> \end{equation} is the Dirac spinor for an antiparticle. Due to crossing symmetry $\hat T$ can be represented in the same way as before (Eq.~(\ref{expans})), with precisely the same functions $c_j$, however in a different $s$, $t$ domain obtained by replacing $p_1'$ by $-p_1'$ and $p_2$ by $-p_2$. The functions $c_j$ arising from (correlated) $\pi\rho$ exchange are assumed to fulfill a dispersion relation over the unitarity cut, \begin{equation} c_j(t,s)=\frac{1}{\pi}\int^\infty_{(m_\pi+m_\rho)^2} \frac{ {\rm Im}\, c_j(t',s)} {t'-t-i\epsilon}dt' \ . \label{Disp} \end{equation} \noindent (Throughout we take the $\rho$ to be a stable particle with $m_\rho$=769 MeV.) Thus the $c_j$ can be determined if their imaginary part is known in the pseudophysical region ($t'\ge (m_\pi+m_\rho)^2$) of the $t$-channel reaction and for $s\ge 4m_N^2$. \subsection{Determination of the spectral functions from unitarity} The required information about the spectral functions $\rho_j(t',s)\equiv {\rm Im}\ c_j(t',s)$ can be obtained from the relevant unitarity relation (cf.\ Fig.~\ref{figtwo}) \begin{equation} i<N\overline N |\hat T-\hat T^\dagger|N \overline N> =\sum_{\pi\rho} \Omega_{\pi\rho}<N\overline N |t^\dagger|\pi\rho> <\pi\rho |t|N \overline N> \delta^{(4)}(k_1+k_2+p_1'-p_1) \label{unit} \end{equation} \noindent where $\Omega_{\pi\rho}$ is a $\pi\rho$ phase-space factor. We first do a partial wave decomposition, \begin{eqnarray} T_t ({\bf p'} \lambda_N'\lambda_{\overline N}'; {\bf p} \lambda_N\lambda_{\overline N};\sqrt t) &=&\frac{1}{4\pi}\sum_J\,(2J+1) d^J_{\lambda\lambda'}(\cos\vartheta) T_t^J ( p' \lambda_N'\lambda_{\overline N}'; p \lambda_N\lambda_{\overline N};\sqrt t)\nonumber \\ t({\bf k} \lambda_\rho; {\bf p} \lambda_N\lambda_{\overline N};\sqrt t) &=&\frac{1}{4\pi}\sum_J\,(2J+1) d^J_{\lambda\bar\lambda}(\cos\bar\vartheta) t^J(k \lambda_\rho; p \lambda_N\lambda_{\overline N};\sqrt t) \label{PWZ} \end{eqnarray} with \begin{equation} \lambda'\equiv \lambda_N'-\lambda_{\overline N}',\quad \bar\lambda\equiv \lambda_\rho-\lambda_\pi=\lambda_\rho, \quad \lambda\equiv \lambda_N-\lambda_{\overline N}\quad, \end{equation} ${\bf p}$, ${\bf p'}$, ${\bf k}$ being the relative 3-momenta in the center-of-mass (cm) system and the angles $\vartheta=\angle({\bf p},{\bf p'})$, $\bar\vartheta=\angle({\bf p},{\bf k})$. After transformation into LSJ basis Eq.~(\ref{unit}) goes into \begin{eqnarray} &&{\rm Im}[T_t^J (p_0,L'\,S';p_0,L\,S;\sqrt t)] \nonumber \\ &&= - C\sum_{L_{\pi\rho}}[t^J(k_0,L_{\pi\rho}\,1;p_0,L'\,S';\sqrt t)] ^\dagger\ t^J(k_0,L_{\pi\rho}\,1;p_0,L\,S;\sqrt t) \equiv\ ^J\!N(L'S';LS) \label{UniRel} \end{eqnarray} \noindent where $C\equiv k_0/(32\pi^2\sqrt t)$ and $p_0$, $k_0$ denote the on-shell momenta of the $N\overline N$ and $\pi\rho$ system, respectively. As for the $2\pi$-exchange case, we want to restrict ourselves to the $J=0,1$ $\pi\rho$ exchange contributions, which act in channels corresponding to the quantum numbers of the pion ($J=0$), and the $\omega$, $A_1$, and $H_1$ meson ($J=1$). Table~\ref{Tabone} shows the quantum numbers and possible transitions from the $N\overline N$ to the $\pi\rho$ system obtained from the conditions \begin{eqnarray} (-1)^{L_{N\overline N}+1}\ \ \ \ \ \ &=&P_\pi P_\rho(-1)^{L_{\pi\rho}} =(-1)^{L_{\pi\rho}}\\ (-1)^{L_{N\overline N}+S_{N\overline N}+I}&=&G_\pi G_\rho =-1 \end{eqnarray} \noindent due to parity and G-parity conservation. We therefore obtain for the different relevant channels \begin{equation} \begin{array}{l} ^0N(00;00) = -C (t^0_{0+})^\dagger t^0_{0+} \hspace{2.6cm} \pi \\ \\ \left. \begin{array}{l} ^1N(01;01) = -C (t^1_{-1})^\dagger t^1_{-1} \hspace{1cm} \\ ^1N(21;01) = -C (t^1_{+1})^\dagger t^1_{-1} \hspace{1cm} \\ ^1N(01;21) = -C (t^1_{-1})^\dagger t^1_{+1} \hspace{1cm} \\ ^1N(21;21) = -C (t^1_{+1})^\dagger t^1_{+1} \hspace{2cm} \\ \end{array} \right \} \omega \\ \\ ^1N(11;11) = -C [(t^1_{1-})^\dagger t^1_{1-}+(t^1_{1+})^\dagger t^1_{1+}] \ \ A_1 \\ ^1N(10;10) = -C [(t^1_{0-})^\dagger t^1_{0-}+(t^1_{0+})^\dagger t^1_{0+}] \ \ H_1 \\ \end{array} \label{LSJamps} \end{equation} The knowledge of $^JN$ determines the spectral functions. With the help of Appendix A, we have explicitly \begin{eqnarray} &&\rho^\pi_P(s,t)\ = \ -{1\over 8\pi\beta^2 }\ \ ^0N(00;00) \nonumber \\ &&\rho^\pi_{\{S,V,T,A\}}(s,t)= 0 \nonumber \end{eqnarray} \begin{eqnarray} \rho^\omega_S(s,t) &=&\ \, \ {1\over 16\alpha^2 \pi\beta}\cos\vartheta\nonumber \\ && \hspace{0.9cm} [\sqrt 2 (\beta-1)^2\ ^1N(01;21) + (2\beta^2+5\beta+2)\ ^1N(21;21) -2(\beta-1)^2\ ^1N(01;01) ] \nonumber \\ \rho^\omega_V(s,t)& =& \ - {1\over 16\alpha^2 \pi\beta}\ [\sqrt 2 (2\beta+1)\ ^1N(01;21) + (\beta^2+2)\; ^1N(21;21) +2(\beta-1)\ ^1N(01;01)] \nonumber \\ \rho^\omega_T(s,t)&=& {1\over 128\,\alpha^2 \pi\beta^2\,m_N^2}\,t\, \nonumber \\ && \hspace{0.9cm} [\sqrt 2 (\beta+2)\ ^1N(01;21) + (2\beta+1)\ ^1N(21;21) -2(\beta-1)\ ^1N(01;01)] \nonumber \\ \rho^\omega_P(s,t)& =& {1\over 16\pi\beta^2} \cos\vartheta [\sqrt 2 (\beta+2)\ ^1N(01;21) + (2\beta+1)\ ^1N(21;21) -2(\beta-1)\ ^1N(01;01)] \nonumber \\ \rho^\omega_{A}(s,t)& =&\ \, \ 0 \nonumber \end{eqnarray} \begin{eqnarray} &&\rho^{A_1}_P(s,t)\ = \ -{3\over 16\alpha^2\pi\beta^2}\ \ ^1N(11;11) \nonumber \\ &&\rho^{A_1}_A(s,t)\ =\ -{3\over 16\alpha^2\pi}\ \ ^1N(11;11)\nonumber \\ &&\rho^{A_1}_{\{S,V,T\}}(s,t) = 0 \nonumber \\ \nonumber \\ &&\rho^{H_1}_P(s,t)\ = \ -{3\over 8\pi\beta^2}\ \cos\vartheta \ \ ^1N(10;10) \nonumber \\ &&\rho^{H_1}_{\{S,V,T,A\}}(s,t) = 0 \\ \label{SpecFunc} \nonumber \end{eqnarray} \noindent with $\beta^2\equiv E^2(p)/m_N^2$ and $\alpha^2\equiv \beta^2-1$. \subsection{Microscopic model for the $N\overline N\to\pi\rho$ process} The determination of the spectral functions $\rho_i$ requires the knowledge of the transition amplitude $t_{N\overline N\to\pi\rho}$ including $\pi\rho$ correlations. In our dynamical model whose structure is visualized in Fig.~\ref{figVPVGT} this quantity is obtained from \begin{equation} t_{N\overline N\to\pi\rho}=v_{N\overline N\to\pi\rho}+ \,v_{N\overline N\to\pi\rho} G_{\pi\rho}T_{\pi\rho\to\pi\rho}\ , \label{nnprlse} \end{equation} \noindent where $v_{N\overline N\to\pi\rho}$ is the transition potential specified later, $G_{\pi\rho}$ is chosen to be the Blankenbecler-Sugar \cite{Blankenbecler66} propagator of the $\pi\rho$ system, and $T_{\pi\rho\to\pi\rho}$ is the $\pi\rho\to\pi\rho$ amplitude essentially taken from the dynamical model \cite{JanssenPRC94}. After partial wave decomposition Eq.~(\ref{nnprlse}) reads more explicitly, in the helicity state basis, \begin{eqnarray} &&t^J(k \lambda_\rho; p \lambda_N\lambda_{\overline N};\sqrt t) = v^J(k \lambda_\rho; p \lambda_N\lambda_{\overline N};\sqrt t) \nonumber \\ && \ \ + \sum_{\lambda_\rho'}\int_0^\infty dk' k'^2 {\omega_\rho(k')+\omega_\pi(k')\over (2\pi)^32\,\omega_\rho(k')\omega_\pi(k')} \ {v^J(k' \lambda_\rho'; p \lambda_N\lambda_{\overline N};\sqrt t)\; T^J(k \lambda_\rho;k'\lambda_\rho';\sqrt t) \over t-(\omega_\rho(k')+\omega_\pi(k'))^2} \quad . \end{eqnarray} \subsubsection{The transition potential $v_{N\overline N\to\pi\rho}$.} Our model for the transition potential $v_{N\overline N\to\pi\rho}$ is based on nucleon and $\Delta$ exchange, together with an $\omega$ pole term (Fig.~\ref{fignnprpot}). In principle, further pole terms exist in the channels considered in this work which are, however, not included in the present calculations for the following reasons: In case of the pion, its mass lies far below the $\pi\rho$ threshold so that such a diagram has a negligible influence on the dispersion integral, Eq.~(\ref{Disp}). In case of the $A_1$ and $H_1$ little is known about their coupling strength to the nucleon. On the other hand, the (bare) $\omega NN$ coupling constant can be fixed by adjusting the final result to the empirical $NN$ repulsion (see below). The starting point for the evaluation of the corresponding potential expressions is the set of interaction Lagrangians \begin{eqnarray} {\cal L}_{\pi NN}&=&\frac{f_{\pi NN}}{m_\pi}\overline\psi\gamma^5 \gamma^\mu\vec\tau\partial_\mu\vec\pi\psi \nonumber \\ \noalign{\vskip5pt} {\cal L}_{\rho NN}&=&g_{\rho NN}\ \overline \psi\left[\gamma^\mu\vec \tau\psi\vec\rho_\mu +\frac{\kappa}{4m_N}\sigma^{\mu\nu}\vec\tau(\partial_\mu\vec \rho_\nu-\partial_\nu \vec\rho_\mu)\right]\psi \nonumber \\ \noalign{\vskip5pt} {\cal L}_{\pi \Delta N}&=&\frac{f_{\pi \Delta N}}{m_\pi}\overline\psi \vec {\cal T}\psi_\mu\partial^\mu\vec \pi \ +\ h.c. \nonumber \\ \noalign{\vskip5pt} {\cal L}_{\rho \Delta N}&=&\frac{f_{\rho \Delta N}}{m_\rho}\overline\psi i \gamma^5\gamma^\mu \vec {\cal T}\psi^\nu (\partial_\mu\vec \rho_\nu-\partial_\nu \vec \rho_\mu)\ +\ h.c. \nonumber \\ \nonumber \\ \noalign{\vskip5pt} {\cal L}_{\omega NN}&=&g_{\omega NN}\ \overline \psi\gamma^\mu\omega_\mu\psi\nonumber \\ \noalign{\vskip5pt} {\cal L}_{\omega\pi\rho}&=&g_{\omega\pi\rho}\epsilon^{\mu\nu\sigma\tau} \partial_\mu\omega_\nu\partial_\sigma\vec\rho_\tau\vec\pi \label{Lagr} \end{eqnarray} We then obtain for the structure of the potential matrix elements \medskip \noindent nucleon exchange: \begin{equation} v_s= i\;f\,F^2\,\frac{f_{NN\pi}g_{NN\rho}}{m_\pi}\ \frac{\bar v(q_2) \left\{ \gamma^5 k\hspace{-5pt}/ _2\ [p\hspace{-5pt}/_x+m_N]\ (\epsilon\hspace{-5pt}/^*- \frac{\kappa}{4m_N}k\hspace{-5pt}/_1 \epsilon\hspace{-5pt}/^* +\frac{\kappa}{4m_N}\epsilon\hspace{-5pt}/^* k\hspace{-5pt}/_1 ) \right\} u(q_1)}{p_x^2-m_N^2} \end{equation} \noindent $\Delta$ exchange: \begin{eqnarray} &&v_s= -i\;f\,F^2\,\frac{f_{N\Delta\pi}g_{N\Delta\rho}}{m_\pi m_\rho}\ \bar v(q_2) \ (k_2)_\mu\ S^{\mu\nu}\ \gamma^5\gamma^\sigma \ [(k_1)_\nu\epsilon^*_\sigma-(k_1)_\sigma\epsilon^*_\nu]\ u(q_1) \nonumber \\ &&\ S^{\mu\nu}=\frac{p\hspace{-5pt}/_x+m_\Delta}{p_x^2-m_\Delta^2} \left\{-g^{\mu\nu}+\frac{1}{3}\gamma^\mu\gamma^\nu+\frac{2}{3m_\Delta^2} p_x^\mu p_x^\nu-\frac{1}{3m_\Delta}(p_x^\mu p_x^\nu-p_x^\nu p_x^\mu)\right\} \ \ \nonumber \\ \end{eqnarray} \noindent $\omega$ exchange: \begin{equation} v_t=-f\,F^2\,\frac{g^{(0)}_{\pi\rho\omega}g^{(0)}_{NN\omega}} {m_\omega}\ \frac{\sqrt t}{p_x^2-(m^{(0)}_\omega)^2} \ \epsilon^{0\nu\sigma\tau}\,(k_1)_\sigma\,\epsilon^*_\tau\ \bar v(q_2)\,\gamma_\nu\,u(q_1) \end{equation} \noindent where $k_1$ ($k_2$) and $q_1$ ($q_2$) denote the four-momenta of the $\rho$($\pi$) and nucleon (antinucleon), respectively. $p_x$ is the momentum of the exchanged particle; for the $\omega$ exchange term, in the cm system, $p_x=(\sqrt t, 0)$. $\epsilon^*$ is the polarization vector for the outgoing $\rho$ meson. $F^2$ denotes the product of vertex form factors, for which we used \begin{eqnarray} s{\rm-channel:}\ \ \ &&F^2=\left ( \frac{2\Lambda_{\pi NX}^2-M^2_X}{2\Lambda_{\pi NX}^2-p_x^2} \right )^2 \left ( \frac{2\Lambda_{\rho NX}^2-M^2_X}{2\Lambda_{\pi NX}^2-p_x^2} \right )^2 \nonumber \\ t{\rm-channel:} \ \ \ \ &&F^2=\left( \frac{\Lambda_{\pi\rho X}^2+m^2_X} {\Lambda_{\pi\rho X}^2+[\omega_\pi(k)+\omega_\rho(k)]^2} \right )^2 \left( \frac{2\Lambda_{NN\, X}^2+m^2_X}{2\Lambda_{NN\, X}^2+4E(p_x)^2} \right )^2 \ \ , \nonumber \\ \label{ffnnpr} \end{eqnarray} \noindent where $X$ stands for the exchanged particle. $f$ denotes the isospin factor; corresponding values are given in Table~\ref{tabparaI}. The coupling constants are either experimentally known or fixed from our former studies. An exception is the bare $\omega NN$ coupling $g^{(0)}_{\omega NN}$ which, as mentioned before, will be fixed later. Values for coupling constants and cutoff masses used are given in Table~\ref{tabparaII}. [The $s$-channel cutoff masses have been adjusted to reproduce the overall strength of $N\overline N \to\pi\rho$ potential used in earlier studies~\cite{Mull91}. This potential produces good agreement with empirical information above the $N\overline N$ threshold, but is based on a different off-shell behavior (time-ordered perturbation theory rather than BbS).] The zero-th components of momenta are determined by the BbS reduction \cite{Aaron76} to be $q_1^0=q_2^0=\sqrt t/2$, $k_1^0=\frac{1}{2}[\sqrt t+\omega_\rho(k)-\omega_\pi(k)]$, and $k_2^0=\frac{1}{2}[\sqrt t+\omega_\pi(k)-\omega_\rho(k)]$. The potential matrix elements are then decomposed into LSJ partial waves in the standard way. Further $u$-channel diagrams arising from $N$ and $\Delta$ exchange can be taken into account by adding a factor of 2 in the (partial wave) $s$-channel contributions. \subsubsection{The amplitude $T_{\pi\rho\to\pi\rho}$.} Our model for the correlation amplitude $T_{\pi\rho\to\pi\rho}$ \cite{JanssenPRC94} is generated from the three-dimensional BbS \cite{Blankenbecler66} scattering equation \begin{equation} T({\bf k'},{\bf k};E)=V({\bf k'},{\bf k};E)+ \int\,d\,^3\,k''\, V({\bf k'},{\bf k''};E) G({\bf k''};E)T({\bf k''},{\bf k};E) \ , \label{LSE} \end{equation} \noindent ($\bf k$, $\bf k'$, $\bf k''$ are corresponding cm relative momenta) with the potential $V$ containing the diagrams shown in Fig.~\ref{figformI}. It contains, besides non-pole pieces, pole terms with bare parameters (masses, coupling constants) which are renormalized by the iteration in Eq.~(\ref{LSE}) and in this way acquire their physical properties. Basic interaction Lagrangians have been taken from the nonlinear $\sigma$-model in the meson sector where the vector mesons are introduced as gauge bosons of a hidden $SU(2)$ or $SU(3)$ symmetry. In this way one obtains the coupling of the $\rho$ to the $\pi$ meson and to itself, i.e.\ ${\cal L}_{\pi\pi\rho}$ and ${\cal L}_{\rho\rho\rho}$ , with a unified value for the coupling constants. Note that we have left out a corresponding pion pole term. The reason is the very small pion mass lying far below the $\pi\rho$ threshold, so that such a diagram should have a negligible influence in the dispersion integral, Eq.~(\ref{Disp}). In addition to the model presented in Ref. \cite{JanssenPRC94} the present calculations include the $H_1$ ($J^P=1^+$, $I^G=0^-$) channel; therefore $V$ now contains an $H_1$-pole term. The corresponding expression is analogous to the $A_1$-term, see \cite{JanssenPRC94}, with the isospin factor $f=3\delta_{I,0}$, bare coupling constant $(g^{(0)}_{H_1 \pi\rho})^2/4\pi=1.3$ and bare mass $m^{(0)}_{H_1}$ = 1100 MeV. As Fig.~\ref{figformII} shows, we obtain a reasonable description of the $H_1$ mass distribution, although compared to the $A_1$ case the model underestimates the empirical situation somewhat. Still, the rough agreement should be sufficient to estimate the relevance of the $H_1$ channel for the correlated $\pi\rho$ exchange $NN$ interaction. \subsection{$NN$ interaction arising from correlated $\pi\rho$ exchange} \label{sect:NNpot} In the last section we specified the dynamical model for the $N\overline N \to\pi\rho$ amplitude, which yields the spectral functions $\rho(s,t)$, Eq.~(\ref{SpecFunc}). The dispersion integral, Eq.~(\ref{Disp}), then determines the invariant amplitudes $c_j(t,s;t<0)$ and thus the scattering operator $\hat T$ (Eq.~(\ref{expans})). The various $NN$ potential contributions are then obtained by sandwiching $\hat T$ between in- and outgoing spinors (cf.\ Eq.~(\ref{invampI})). Such a calculation can be directly pursued for the $\pi$ and $A_1$ channel since these spectral functions do not depend on $\cos\vartheta$, which is, in terms of the Mandelstam variables, \begin{equation} \cos\vartheta \ = \ {4m_N^2-t-2s \over t-4m_N^2}\ = \ {u-s \over t-4m_N^2} \ \ . \label{cos} \end{equation} When transforming into the $NN$ channel, corresponding values for $s>4 m_N^2$ have to be inserted, and, in principle, the $t$-dependence in $\cos\vartheta$ should be integrated over in the dispersion integral. However, it is then not guaranteed that the typical structure of $s$-channel vector meson exchange is obtained. Namely, starting from the conventional vector meson Lagrangian, \begin{equation} {\cal L}\ = \ g\, \bar \psi\gamma^\mu\psi\, V_\mu \;+\; f/4m_N\;\bar \psi \sigma^{\mu\nu}\psi\,(\partial_\mu\,V_\nu -\partial_\nu\,V_\mu)\quad, \end{equation} \noindent one obtains for the scattering operator ($m_V$ is the mass of the vector meson) \begin{equation} {\hat T} \, = \, {1\over t-m_V^2} \left\{g^2\,[-V]-gf\,[\,V+{u-s\over 4m_N^2}S+{t\over 8m_N^2}T +{u-s\over 4m_N^2}P\,]-f^2\,[\,{t\over 8m_N^2} T + {u-s\over 4m_N^2}P\,]\right\} \ \ . \label{Vectexneu} \end{equation} Obviously a characteristic factor $u-s \sim \cos\vartheta$ occurs in front of the invariants $S$ and $P$ as well as a factor $t$ in front of $T$. This structure is, for the case of $\rho$ exchange, of decisive importance for a correct behavior of the $NN$ interaction. However, since we have to apply approximations when evaluating the dispersion integral (by introducing a cutoff $t_c = 4m_N^2$), this structure is not automatically obtained when doing a straightforward calculation. Therefore we decided to transform these factors directly into the $NN$ channel and to apply the dispersion integral for the remaining part of the spectral functions. (In a more formal language, new invariant operators have to be defined which include these factors.) Trivially the results are then forced to have the structure of $s$-channel vector meson exchange. Another important modification of the above formulas remains to be introduced. So far, by construction (see Fig.~\ref{figVPVGT}), our results contain not only the correlated part we are interested in, but also the uncorrelated contribution, cf.\ Fig.~\ref{figintI}. Therefore the latter has to be removed. We do this by subtracting the Born term part of the $^JN$ functions of Eq.~(\ref{LSJamps}), which leads to new functions $^JN_{corr}$ given by \begin{equation} ^JN_{corr}(00;00)=-C[(t^0_{0+})^\dagger t^0_{0+}-(v^0_{0+})^\dagger v^0_{0+}] \end{equation} \noindent for the pion and analogous extensions for the other channels. These new functions $^JN_{corr}$ are actually used when evaluating the spectral functions by means of Eq.~(\ref{SpecFunc}). We still have to transform the isospin part of the $N\overline N\to N\overline N$ amplitude into the $s$-channel. Resulting isospin factors are provided by the isospin crossing matrix~\cite{Martin70}. In general we have \begin{eqnarray} &&f^{I=0}_{NN} = {1\over 2}(f^{I=0}_{N\overline N} -3f^{I=1}_{N\overline N}) \nonumber \\ &&f^{I=1}_{NN} = {1\over 2}(f^{I=0}_{N\overline N} +f^{I=1}_{N\overline N}) \end{eqnarray} \noindent for the connection of isospin factors in both channels. The factors for our $N\overline N\to N\overline N$ amplitude are already implicitly taken into account by including corresponding factors in $v_{N\overline N \to \pi\rho}$ and $T_{\pi\rho\to\pi\rho}$. For the (isospin zero) $\omega$ and $H_1$ channels we thus have $f^I_{N\overline N}=\delta_{I,0}$ whereas in $\pi$ and $A_1$ we have $f^I_{N\overline N}=\delta_{I,1}$. Therefore our final result for the correlated $\pi\rho$ exchange $NN$ potential can be written as operator in isospin space in the following way: \begin{equation} V_{\pi\rho,corr}=\frac{\kappa}{2}\sum_i \left [ \sum_{\alpha=\omega,H_1} \int^{t_c}_{(m_\pi+m_\rho)^2}dt' \frac{\rho^\alpha_i(t')}{t'-t} C_i {\bf 1} +\sum_{\beta=\pi,A_1} \int^{t_c}_{(m_\pi+m_\rho)^2}dt' \frac{\rho^\beta_i(t')}{t'-t} C_i {\bf \tau}_1\cdot {\bf \tau}_2 \right ] \quad, \end{equation} \noindent where $C_i$, according to the foregoing discussion, are matrix elements between nucleon helicity spinors of slightly modified invariants $(u-s)S$, $(u-s)P$, $V$, $A$, $tT$. The factor $\kappa=\frac{1}{(2\pi)^3} \frac{m_N^2}{\sqrt{E_1E_1'E_2E_2'}}$ arises because $V_{\pi\rho,corr}$ is to be defined as part of the Bonn potential whose $T$-matrix is defined by \begin{equation} S_{fi}=\delta_{fi}-i\,2\pi\,\delta^{(4)}(p_1'+p_2'-p_1-p_2) T_{fi} \ . \end{equation} In order to be used in a scattering equation the resulting potential must be given off shell, as function of the in- and outgoing cm relative momenta and total energy in the $s$ ($NN$) channel, i.e.\ $V = V(p',p;E_{cm})$. On shell, for $p'=p=p_0$ with $E_{cm}^2 = 4(m_N^2+p_0^2)$, the relation of these variables to the Mandelstam variables is unique and given by \begin{eqnarray} s&=&4E(p)^2 \nonumber \\ t&=&-2p^2(1-\cos\vartheta) \nonumber \\ u&=&-2p^2(1+\cos\vartheta) \ . \label{on} \end{eqnarray} Half-off-shell, i.e.\ for $p'\ne p$, we take the plausible prescription $t=-({\bf p}-{\bf p'})^2$ and $s=4E(p)E(p')$, which of course agrees with Eq.~(\ref{on}) on shell. Dependence on the starting energy is assumed to be of the same type as in time-ordered perturbation theory applied in the Bonn potential. Here the propagator of an exchanged meson reads \begin{equation} \frac{1}{\omega\, (E_{cm}-E(p)-E(p')-\omega)} \ . \vspace{0.3cm} \end{equation} In order to obtain a natural generalization of this expression we first define the `on-mass-shell energy of an exchanged $\pi\rho$ system', $\Omega\equiv\sqrt{t'+({\bf p}-{\bf p'})^2}=\sqrt{t'-t}$ and replace the energy denominator of the dispersion integral by the on-shell-equivalent expression \begin{equation} \frac{1}{t'-t}\to\frac{1}{\Omega\, (E_{cm}-E(p)-E(p')-\Omega)} \ . \vspace{0.3cm} \end{equation} Finally the resulting potentials have to fall off sufficiently rapidly in order to be able to solve the scattering equation. For this reason we introduce an additional form factor, $\left (\frac{n\Lambda^2-m^2}{n\Lambda^2-t}\right )^n \to\left (\frac{n\Lambda^2-t'}{n\Lambda^2-t} \right )^n$, with $\Lambda=$5 GeV, $n$=5, into the dispersion integral. The large cutoff mass chosen ensures that the results are not modified on shell. \subsection{Determination of effective coupling constants and masses} \label{sect:effcoup} It is convenient to parametrize our correlated $\pi\rho$ exchange results in terms of single, sharp-mass effective meson exchange, as done for the analogous case of correlated $\pi\pi$ in the $NN$ \cite{Kim94} and $\pi N$ \cite{Schuetz94} system. For reasons to be discussed below, this can only be done successfully for the $\pi$ and $\omega$ channel contributions. For an effective $\pi'$ the scattering operator reads \begin{equation} \hat T= -g_{\pi'NN}^2\, \vec \tau_1\vec\tau_2\, \frac{1} {t-m_{\pi'}^2}\, P \end{equation} For the $\omega'$, the expression has already been given in Eq.~(\ref{Vectexneu}). By comparison of the coefficients belonging to the invariants with the corresponding dispersion-theoretic terms one can determine an effective coupling constant, which will in general be $t$-dependent. For example we have for the pionic channel \begin{equation} -g_{\pi'NN}^2(t)\;\frac{1}{t-m_{\pi'}^2}\ = \ \frac{1}{2} \left [ {1\over\pi}\int\limits_{(m_\pi+m_\rho)^2}^\infty {\rho^\pi_P(t')\over t'-t} dt'\right ]\ \ , \end{equation} The effective mass $m_{\pi'}$ is chosen such that $g_{\pi' NN}$ becomes essentially independent of $t$. It turns out that for the $H_1$ channel such a mass cannot be found. For channels involving several invariants, different coupling constants (and masses) are obtained which depend on the specific invariant for which the comparison is made. Obviously such a parametrization is successful if the resulting values only weakly (if at all) depend on the invariant chosen. This is the case for the $\omega$-channel but not for the $A_1$-channel. \section{Results and discussion} \label{sect:results} Having introduced the necessary formalism for the evaluation of correlated $\pi\rho$ exchange we now present the results and investigate their consequences for the $NN$ interaction. We start with the results obtained in the $t$-channel, i.e.\ for the $N\overline N \to\pi\rho$ amplitude and the spectral functions. We then discuss the properties of the resulting potential contribution, which is obtained from the (dispersion-theoretic) transformation into the $s$-channel, and point out its role within the Bonn meson exchange $NN$ interaction \cite{Machleidt87}. \subsection{The $\pi$ channel} The pionic channel of correlated $\pi\rho$ exchange is of special importance. As mentioned in the introduction it is a natural candidate to provide additional (short ranged) tensor force required to fit empirical $NN$ data with models using a realistic, soft $\pi NN$ form factor. Fig.~\ref{figresI} shows the $N\overline N \to\pi\rho$ (on-shell) potential $v_{N\overline N \to\pi\rho}$ and the corresponding amplitude $t_{N\overline N \to\pi\rho}$ obtained from Eq.~(\ref{nnprlse}), which contains the effect of $\pi\rho$ correlations. The purely imaginary potential has a strong increase at the pseudophysical $\pi\rho$ threshold and a maximum near $t=50m_\pi^2$. $\pi\rho$ correlations strengthen this maximum; in addition they generate a real part in the amplitude. Note that these modifications act quadratically in the $NN\to NN$ amplitude and therefore also in the $NN$ potential, so that the final effect of $\pi\rho$ correlations is stronger than Fig.~\ref{figresI} suggests. Indeed, the spectral function $\rho_P^\pi(t)$ in Fig.~\ref{figresII} demonstrates that the piece due to correlated $\pi\rho$ exchange has a considerable strength, although it is smaller than the uncorrelated part generated by the transition potential only. One has to keep in mind that a good part of the latter contribution consists of iterative $\pi\rho$ box diagrams involving $NN$ intermediate states, which are not part of $V_{NN}$ but are generated by the $NN$ scattering equation. Consequently the role of uncorrelated processes in $V_{NN}$ is considerably smaller than suggested from the figure. The spectral function of the correlated part has a clear maximum at about $t$=60 $m_\pi^2$, thus representing the mass distribution of a broad, heavy effective particle with pionic quantum numbers. A rough parametrization of this contribution by sharp-mass particle exchange appears to require a mass of about 1 GeV, noticeably smaller than chosen in Ref. \cite{Holinde90} for the phenomenological $\pi'$. In the $s$-channel we first demonstrate the influence of the resulting on-shell potentials in the $^3S_1-\,^3D_1$ and $^1S_0$ partial waves as function of the nucleon lab energy. In Fig.~\ref{figresIII} the dotted and dashed curves show the corresponding one-pion-exchange (OPE) potentials; obviously there is a strong suppression of OPEP when the $\pi NN$ cutoff mass is reduced from the value of 1.3 GeV used in the full Bonn potential to 1 GeV. Most importantly, the addition of the potential due to correlated $\pi\rho$ exchange in the pionic channel to the dotted curve restores the original tensor force strength; obviously it is able to counterbalance the suppression induced by the smaller $\pi NN$ cutoff mass. The $^3S_1-\,^3D_1$ partial wave is of special importance in this connection since it exclusively contains a tensor force component, which is decisive for a realistic description of deuteron properties. It should be added that the above result is in remarkable agreement with a previous calculation of the $\pi NN$ form factor \cite{JanssenPRL94}, Fig.~\ref{figresIV}, which consistently used the same $\pi\rho$ $t$-matrix and $N\overline N \to\pi\rho$ transition potential, and independently arrived at $\Lambda_{\pi NN}$ = 1 GeV. As discussed in Sect.~\ref{sect:effcoup}, correlated $\pi\rho$ exchange in the $\pi$ and $\omega$ channels can be parametrized by sharp-mass one-boson-exchange (OBE) potentials, provided that the effective coupling constants are allowed to become $t$-dependent. Fig.~\ref{figresV} shows such coupling constants for the exchange of a heavy $\pi'$ for various chosen masses of $\pi'$. Obviously the coupling becomes $t$-independent for a mass of 1020 MeV, quite near the maximum of the correlated spectral function in Fig.~\ref{figresII}, and the resulting coupling constant is $g_{\pi' NN}^2/4\pi\simeq 9$. If we choose $m_{\pi'}$ =1200 MeV as in \cite{Holinde90}, the resulting coupling constant is noticeably $t$-dependent; its strength is much smaller than used in \cite{Holinde90}. There are two reasons for this discrepancy: First the authors of \cite{Holinde90} applied a $\pi' NN$ form factor, which reduces the strength at $t$=0 (relevant for $NN$ scattering) by more than a factor of 2. Second, the strength of the $\pi'$ was phenomenologically chosen in \cite{Holinde90} to compensate for a much softer $\pi NN$ form factor, with a cutoff mass of 800 MeV. (Indeed a much smaller value ($g_{\pi' NN}^2/4\pi$= 70) is sufficient to compensate for a form factor with $\Lambda_{\pi NN}$ = 900 MeV~\cite{Haidenbauer94}.) Obviously correlated $\pi\rho$ exchange can only partly explain the phenomenological $\pi'$; another possible mechanism is correlated $\pi\sigma$ exchange \cite{Ueda92} (where $\sigma$ stands for correlated $\pi\pi$ exchange in the S-wave channel). In order to show that the compensation for a softer $\pi NN$ form factor by correlated $\pi\rho$ exchange is valid not only for on-shell potentials, but also for $NN$ amplitudes and observables, we extrapolate the correlated $\pi\rho$ exchange off shell as described in Sect.~\ref{sect:NNpot}, add this piece to the (full) Bonn potential (with a reduced $\pi NN$ cutoff mass of 1 GeV) and solve the relativistic Schroedinger equation relevant for the Bonn potential. Only a slight readjustment of the coupling of the isospin-one scalar meson ($\delta$ in \cite{Machleidt87}) in the original Bonn potential is required in order to obtain again a good description of $NN$ phase shifts. As Table~\ref{TABdeut} demonstrates convincingly the deuteron observables also can be reproduced with a considerably softer $\pi NN$ form factor (characterized by $\Lambda_{\pi NN}$ = 1 GeV), provided that correlated $\pi\rho$ exchange in the pionic channel is included. \subsection{The $\omega$-channel} Since the mass of the physical $\omega$-meson is only slightly below the $\pi\rho$ threshold, genuine pole terms have been included in the $\pi\rho$ amplitude \cite{JanssenPRC94} as well as in our model for the transition potential $v_{N\overline N\to \pi\rho}$. As Fig.~\ref{figresVI}(a) shows, the contribution of such pole terms leads to a reduction of the (imaginary) transition potential above the $\pi\rho$ threshold, whose amount depends on the value of the bare coupling constant $g^{(0)}_{\omega NN}$. (The reason for our choice $g^{(0)}_{\omega NN}/4\pi$ = 4.40 will be discussed later.) Fig.~\ref{figresVI}(b) shows the resulting amplitude $t_{N\overline N\to \pi\rho}$. Similarly to the pionic channel, $\pi\rho$ correlations enhance the maximum of the amplitude near threshold. The inclusion of the $\omega$-meson pole terms in our dynamical model ensures that the imaginary part of the $N\overline N\to N\overline N$ amplitude, and therefore the resulting $NN$ potential, contains, besides `true' correlated $\pi\rho$ exchange (Fig.~\ref{figresVII}(e)) generated by the non-pole parts of the corresponding amplitudes, also genuine $\omega$ exchange processes (Fig.~\ref{figresVII}(a)-(d)). Corresponding propagators and vertex functions are dressed by $\pi\rho$ loop corrections. For example, the $\omega$ propagator has the following structure (for details we refer the reader to \cite{JanssenPRC94}): \begin{equation} d=\frac{1}{t-(m^{(0)}_{\omega})^2-\Sigma(t)}, \quad {\rm with}\quad \Sigma(t)\sim \int f^{(0)} G_{\pi\rho} f \quad, \label{prop} \end{equation} \noindent where $f^{(0)}$ and $f$ are bare and dressed $\omega\to\pi\rho$ vertex functions, respectively, and $G_{\pi\rho}$ denotes the $\pi\rho$ propagator. The bare parameters $g_{\omega\pi\rho}^{(0)}$ and $m^{(0)}_\omega$ have been adjusted such that $d$ has a pole at $t=m_\omega^2$. The imaginary part of the $N\overline N\to N\overline N$ amplitude, and therefore the corresponding spectral functions, consist of a $\delta$-function at $t=m_\omega^2$, which precisely corresponds to the exchange of a (physical) $\omega$-meson with point-like $\omega NN$ coupling, i.e.\ without any form factor. However it is important to realize that $d$, and therefore the diagrams in Fig.~\ref{figresVII}(a)-(d), provide a further contribution, since $\pi\rho$ intermediate states make the self-energy $\Sigma$ complex above the $\pi\rho$ threshold. Such intermediate states likewise occur at the vertices in diagrams (b)-(d), leading to additional contributions to the spectral functions. Fig.~\ref{figresVIII} shows the resulting spectral functions. Note that although we assumed the bare $\omega NN$ coupling to be of pure vector type, small contributions to $\rho_S$, $\rho_T$, and $\rho_P$ occur, which are generated by the $\pi\rho$ loops in Fig.~\ref{figresVII}(b)-(d). First we have the $\delta$-function piece (dashed); above $\pi\rho$ threshold we have additional contributions from diagrams \ref{figresVII}(a)-(d) (dash-dotted) which have opposite sign to the $\delta$-function. They act as vertex corrections which suppress the point-like coupling of $\omega$ exchange and thus generate a form factor effect. Finally there is a sizable non-pole contribution (dotted curve); throughout it has opposite sign to the part generated by the vertex corrections and roughly counterbalances their effect, a fact found already in the pionic channel. Again, after presenting the results in the $t$-channel, we now want to look at the corresponding on-shell potentials in the $s$-channel. Since we deal with rather short-ranged contributions we show, as two characteristic examples, the results for the $^1S_0$ and $^3P_1$ partial waves (Fig.~\ref{figresIX}). Note that the bare $\omega NN$ coupling ($g^{(0)}_{\omega NN}$=4.40), which determines the size of diagrams (a)-(c), has been chosen such that the total repulsion generated by all diagrams of Fig.~\ref{figresVII} agrees (at low energies) with the effective $\omega$ exchange in the Bonn potential needed empirically. The dashed curves are generated by the $\delta$-functions in Fig.~\ref{figresVIII}, with a predicted renormalized coupling constant of $g^2_{\omega NN}$= 11.0. Apparently this contribution alone provides almost the same repulsion as in the Bonn potential although the coupling constant is about a factor of 2 smaller. The reason is that the phenomenological form factor in the Bonn potential, with the monopole cutoff mass (1.5 GeV) of only twice the $\omega$ mass, leads to a strong reduction of the coupling constant in the physical region ($g^2_{\omega NN}$($t$=0)=10.6). The vertex corrections (generated by diagrams (a)-(d) above $\pi\rho$ threshold) strongly reduce the repulsion in the physical region, leading to the dash-dotted curve. This suppression is essentially cancelled by the `true' correlated $\pi\rho$ exchange (diagram \ref{figresVII} (e)), as already demonstrated for the spectral functions. Obviously the latter contribution is remarkably strong; it explains about 40\% of effective $\omega$ exchange. The new reduced coupling constant (11.0) is still about a factor of 2 larger than provided by customary SU(3) estimates, which use $g^2_{\omega NN}$= 9$g^2_{\rho NN}$. Thus with $g^2_{\rho NN}/4\pi$=0.55 as determined by Hoehler and Pietarinen \cite{Hoehler75} we have $g^2_{\omega NN}/4\pi\simeq$5. Note however that the above relation between $\omega$ and $\rho$ coupling constants is based, apart from ideal mixing, on the assumption of vanishing $\phi NN$ coupling. For $g_{\phi NN}$ unequal to zero the above relation goes into \begin{equation} g_{NN\omega}=3g_{NN\rho}-\sqrt 2 g_{NN\phi} \ . \end{equation} Thus if we take $g_{\phi NN}$= -$g_{\rho NN}$ (which amounts to a rather small deviation from zero) we have $g^2_{\omega NN}\approx 20 g^2_{\rho NN}$, in rough agreement with our results. Such a value for the $\phi NN$ coupling to the nucleon and the negative sign is well conceivable, if the $\phi$ couples to the nucleon via the $K\overline K$ continuum \cite{Mullpriv}. As discussed before, for practical reasons it is convenient to parametrize also the non-pole contribution of diagram \ref{figresVII}(e) by an effective one-boson-exchange. Results are shown in Fig.~\ref{figresX}, for the dominant vector as well as the tensor coupling. Obviously they can be reasonably represented by $g^2_{\omega' NN}/4\pi$=8.5, $f_{\omega' NN}/g_{\omega' NN}$=0.4, and $m_{\omega'}$=1120 MeV. Using the mass of the physical $\omega$ meson $m_\omega=$782 MeV we find a ($t$-dependent) effective $\omega$ coupling strength characterized by $g^2_{\omega NN} \simeq$ 4. It is interesting to note that the suppression of the tensor coupling, in some sense enforced in the pole terms by assuming the corresponding bare coupling to be zero, also happens in the non-pole term. \subsection{The $A_1$/$H_1$-channel} After the discussion of the $\pi$ and $\omega$ channel we now want to investigate the $A_1$ and $H_1$ channels together since their structure is very similar. Compared to the $\pi$ and $\omega$ channels we have important differences. First the $A_1$ as well as the $H_1$ mass lie above the $\pi\rho$ threshold, and both particles decay with a very large width into $\pi\rho$. Consequently their propagators now acquire a pole in the complex plane. Second, in contrast to the $\pi$ and $\omega$ channels, the Bonn potential \cite{Machleidt87} does not contain $A_1$/$H_1$ OBE contributions, which could be used to fix the bare $A_1 NN$ and $H_1 NN$ coupling constants, as was done in the $\omega$ channel. Since there is no other {\em a priori} information about these couplings, we will in this first extrapolatory study, simply put them to zero, i.e.\ take only diagrams of type \ref{figresVII}(d) and (e) into account. If, in a later stage, those couplings turn out to be needed (e.g.\ in order to obtain a quantitative fit to the $NN$ data), they should be included. There is a further structural difference which has an enormous impact on the results and requires an extended discussion. In general, besides the unitarity cut for $t>(m_\pi +m_\rho)^2$ treated in the dispersion integral (Eq.~(\ref{Disp})), there exists a left hand cut for $t<t_0$ in the $N\overline N \to \pi\rho$ amplitude generated by $s$-channel poles due to nucleon and $\Delta$-isobar exchange (Fig.~\ref{fignnprpot}). $t_0$ is fixed by the condition $s-m^2_{N/\Delta}$=0. There are two solutions for each exchange; the largest value (generated by nucleon exchange) is at $t_0\approx 42 m_\pi^2$, i.e.\ just below the $\pi\rho$ threshold. Consequently this branch point will influence the resulting potentials near threshold considerably. In the $\pi$ and $\omega$ channels the corresponding potentials act in P-waves and are thus proportional to the $\pi\rho$ on-shell momentum $k_0$, with the effect that the corresponding transition potentials start to increase first when approaching the threshold, but are then suppressed by the $k_0$ factor. In this way one obtains the characteristic structure of a maximum near threshold, which we have observed in such channels. The point now is that both the $A_1$ and $H_1$ are $\pi\rho$ S-waves; therefore the transition potentials do not contain the damping factor $k_0$ anymore. Indeed, Figs.~\ref{figresXII} and \ref{figresXIII} show the overwhelming effect of the left hand cut near threshold, in both channels. It essentially remains when $\pi\rho$ correlations (which contain the $A_1$ resonance) are included, although the influence of the $A_1$ is clearly seen. Fig.~\ref{figresXIV} shows the resulting spectral functions $\rho_P^{H_1}$, $\rho_A^{A_1}$, and $\rho_P^{A_1}$. Again the strong effect of the left hand cut near threshold is obvious. Note also that for $\rho_P$, $A_1$ and $H_1$ provide roughly similar contributions, but with opposite sign. In Fig.~\ref{figresXV} we present the resulting on-shell potentials in some selected partial waves. For both S-states, the (attractive) $A_1$ contribution strongly dominates the result arising from the $H_1$ channel but is considerably smaller (as far as the modulus is concerned) compared to the corresponding piece in the $\omega$ channel, cf.\ Fig.~\ref{figresX}. For $^3P_1$ both contributions have opposite sign and roughly cancel; the total result is negligible compared to the $\omega$ channel. In contrast to the $\pi$, $\omega$ channels discussed before, the above results cannot be suitably parametrized in terms of sharp-mass exchanges. In case of the $A_1$, no reasonable mass can be found which works for both spectral functions; moreover all effective coupling constants become strongly $t$-dependent. The basic reason is again the dominance of the left hand cut, which destroys the conventional bump structure of the spectral functions found in other channels of correlated $\pi\pi$ and $\pi\rho$ exchange. \section{Concluding remarks} \label{sect:concl} In this work we have determined the contribution to the $NN$ interaction due to the exchange of a correlated $\pi\rho$ pair between two nucleons. The correlations between $\pi$ and $\rho$ have been taken into account using a realistic meson exchange model of the $\pi\rho$ interaction \cite{JanssenPRC94}. In a first step we evaluated the $t$-channel amplitude $N\overline N \to N\overline N$ including $\pi\rho$ correlations. The transformation into the $s$-channel with the help of dispersion-theoretic methods then yields the correlated $\pi\rho$ exchange $NN$ potential. We have investigated four relevant channels of the $\pi\rho$ system characterized by the quantum numbers of the physical particles $\pi$, $\omega$, $A_1$, and $H_1$. In the pionic channel correlated $\pi\rho$ exchange yields a short-ranged contribution, which roughly corresponds to an exchange of a heavy (effective) $\pi'$ with a mass of about 1 GeV. The additional tensor force generated by this potential is sufficient to compensate for a reduction of the $\pi NN$ cutoff mass $\Lambda_{\pi NN}$ from 1.3 GeV to 1.0 GeV in the one-pion-exchange potential. For basic theoretical reasons, such a reduction is highly welcome, since various models of nucleon structure unanimously predict a rather soft $\pi NN$ form factor characterized by $\Lambda_{\pi NN}\simeq 0.8\ {\rm GeV}$. Such a small value might be reached if correlated $\pi\sigma$ exchange is included, too, which is also missing in the Bonn potential. (As usual, $\sigma$ stands for a low mass correlated $\pi\pi$ pair in the $0^+$ channel.) Thus it appears that in the Bonn potential~\cite{Machleidt87} the one-pion-exchange potential together with a hard form factor is an effective description of `true' one-pion exchange (with a soft form factor) plus correlated $\pi\rho$ (and $\pi\sigma$) exchange in the pionic channel. In the $\omega$ channel, the exchange of a correlated $\pi\rho$ pair also provides a sizable contribution to the $NN$ interaction. Since the $\omega$ mass is near the $\pi\rho$ threshold we have included the genuine $\omega$-meson explicitly and replaced the (effective) $\omega$ exchange in the Bonn potential by the resulting correlated $\pi\rho$ potential, which can be decomposed into a pole and a non-pole piece. The former provides a microscopic model for `true' $\omega$ exchange leading to a renormalized $\omega NN$ coupling constant, $g^2_{\omega NN}/4\pi$ = 11.0, which is about a factor of two smaller than the effective value of 20 used in the Bonn potential. Thus `true' correlated $\pi\rho$ exchange (Fig.~\ref{figintI}(e)) provides almost half of the empirical repulsion needed in the $NN$ interaction; it can be parametrized by sharp-mass $\omega'$ exchange with $g^2_{\omega'NN}\simeq 8.5$, $f_{\omega' NN}/g_{\omega' NN}\simeq$ 0.4 and $m_{\omega '}\simeq$ 1120 MeV. Our present result for the $\omega$ coupling constant ($g_{\omega NN}^2\approx 20 g_{\rho NN}^2 $) is well compatible with $SU(3)$, provided that there exists a small, negative $\phi NN$ coupling of vector type, with the magnitude of the order of the $\rho$-coupling. Such a $\phi NN$ coupling (especially the required negative sign) occurs naturally if it is supposed to arise via the $K\overline K$ continuum. Although corresponding $\phi$ exchange in the $NN$ interaction provides only a small contribution to the repulsion, it makes the above relation between $\omega$ and $\rho$ couplings agree with $SU(3)$ predictions. Consequently there appears to be little room for explicit quark-gluon effects in being responsible for the short-range $NN$ repulsion. Additional contributions arise in the $A_1$ and $H_1$ channels. They are sizable individually, mainly due to left-hand cut effects arising from nucleon and $\Delta$ exchange in a $\pi\rho$ S-wave. However, in some partial waves strong cancellations occur between the $A_1$ and $H_1$ contributions. Further contributions are, in principle, generated by direct coupling of the $A_1$/$H_1$ to the nucleon. The size of such couplings is, however, not known; these terms are therefore omitted in the present work. It remains to be seen whether a precise fit of the $NN$ observables requires such terms and thereby establishes their existence. \begin{appendix} \section{Determination of spectral functions} In this appendix we derive Eq.~(\ref{SpecFunc}) of the main text, which provides the connection between the spectral functions needed in Eq.~(\ref{Disp}) and the imaginary part of the $N\overline N\to N\overline N$ amplitude in the LSJ basis ($^JN$) obtained from the unitarity relation (Eq.~(\ref{unit})). For the latter it was necessary to work in the LSJ basis in order to identify the allowed $N\overline N \to \pi\rho$ transitions. In order to establish the connection to the spectral functions however, matrix elements of the invariants $\hat C_j$ are required, which is most suitably done in the helicity state basis. The imaginary part of the $N\overline N\to N \overline N$ amplitude is now defined by (cf.\ Eq.~(\ref{UniRel}) for the analogous definition in LSJ basis) $^JN( \lambda_N' \lambda_{\overline N}';\lambda_N\lambda_{\overline N})\equiv {\rm Im}\,T^J (p' \lambda_N' \lambda_{\overline N}'; p \lambda_N\lambda_{\overline N};\sqrt t)$ and the relation between LSJ and helicity basis amplitudes is given by the standard expressions (cf.\ \cite{Machleidt87}) \begin{eqnarray} ^J N_1&=&\ \frac{1}{2}\ ^JN(J0;J0)+a^2\ ^JN(J-1 1;J-11)-ab\ ^JN(J+11;J-11) \nonumber \\ && \ \ \ \ -ab\ ^JN(J-11;J+11)+b^2 \ ^JN(J+11;J+11) \nonumber \\ ^JN_2&=&-\frac{1}{2}\ ^JN(J0;J0) +a^2\ ^JN(J-1 1;J-11)-ab\ ^JN(J+11;J-11) \nonumber \\ && \ \ \ \ -ab\ ^JN(J-11;J+11)+b^2 \ ^JN(J+11;J+11) \nonumber \\ ^J N_3&=&\ \ \frac{1}{2}\ {^JN(J1;J1)}+b^2\ ^JN(J-1 1;J-11)+ab\ ^JN(J+11;J-11) \nonumber \\ && \ \ \ \ +ab\ ^JN(J-11;J+11)+a^2 \ ^JN(J+11;J+11) \nonumber \\ ^JN_4&=&-\frac{1}{2}\ {^JN(J1;J1)}+b^2\ ^JN(J-1 1;J-11)+ab\ ^JN(J+11;J-11) \nonumber \\ && \ \ \ \ +ab\ ^JN(J-11;J+11)+a^2 \ ^JN(J+11;J+11) \nonumber \\ ^J N_5&=&\ \ ab \ ^JN(J-1 1;J-11)-b^2\ ^JN(J+11;J-11) +a^2\ ^JN(J-11;J+11) \nonumber \\ && \ \ \ \ -ab \ ^JN(J+11;J+11) \nonumber \\ ^JN_6 &=&\ ^JN_5 \ \ \ \ (\rm on-shell) \label{bla} \end{eqnarray} where we used the short-hand notation for the $^JN$ amplitudes defined in Table~\ref{bull} and \begin{equation} a=\sqrt{{J\over 2(2J+1)}} \ \ \ \ \ \ \ b=\sqrt{{J+1 \over 2(2J+1)}} \ \ {}. \end{equation} \noindent The various channel contributions to ${\rm Im} T({\bf p'}\lambda_N'\lambda_{\overline N}'; {\bf p} \lambda_N\lambda_{\overline N};\sqrt t)$ are then given by \begin{eqnarray} \pi:&& N_i\equiv N( \lambda_N' \lambda_{\overline N}';\lambda_N\lambda_{\overline N})\equiv \frac{1}{4\pi}d^0_{\lambda\lambda'}(\cos\vartheta)\ ^0N( \lambda_N' \lambda_{\overline N}';\lambda_N\lambda_{\overline N}) \equiv\frac{1}{4\pi}d^0_{\lambda\lambda'}(\cos\vartheta)\ ^0N_i \nonumber \\ \omega,A_1,H_1:&& N_i\equiv N( \lambda_N' \lambda_{\overline N}';\lambda_N\lambda_{\overline N})\equiv \frac{3}{4\pi}d^1_{\lambda\lambda'}(\cos\vartheta)\ ^1 N(\lambda_N' \lambda_{\overline N}';\lambda_N\lambda_{\overline N}) \equiv \frac{3}{4\pi}d^1_{\lambda\lambda'}(\cos\vartheta)\ ^1N_i \nonumber \\ \end{eqnarray} We now define a vector $ {\bf N}^\alpha=\left ( \begin{array}{ccccc} N_1^\alpha & N_2 ^\alpha& N_3^\alpha & N_4^\alpha & N_5^\alpha \\ \end{array} \right ) $, $\alpha=\pi,\omega,A_1, H_1$, \noindent and use Eqs.~(\ref{LSJamps}) and (\ref{bla}) to express its components in terms of the LSJ amplitudes $^JN(L'S';LS)$ for each of the contributing channels. We obtain \begin{equation} {\bf N}^\pi=\left ( \begin{array}{c} \frac{1}{8\pi}\ ^0N(00;00) \\-\frac{1}{8\pi}\ ^0N(00;00) \\ 0 \\ 0 \\ 0 \end{array} \right ) \quad, \end{equation} \begin{equation} {\bf N}^{A_1}=\left ( \begin{array}{c} 0 \\ 0 \\ \frac{3(1+\cos\vartheta)}{16\pi}\ ^1N(11;11) \\ \frac{3(-1+\cos\vartheta)}{16\pi}\ ^1N(11;11) \\ 0 \end{array} \right ) \quad , \end{equation} \begin{equation} {\bf N}^{H_1}=\left ( \begin{array}{c} \frac{3\cos\vartheta}{8\pi}\ ^1N(00;00) \\ -\frac{3\cos\vartheta}{8\pi}\ ^1N(00;00) \\ 0 \\ 0 \\ 0 \end{array} \right ) , \end{equation} \begin{equation} {\bf N}^\omega=\left ( \begin{array}{c} \frac{3}{4\pi}\cos\vartheta [\frac{1}{6}\ ^1N(01;01) -\sqrt{\frac{2}{9}}\ ^1N(01;21)+\frac{1}{3}\ ^1N(21;21)] \\ \frac{3}{4\pi}[\frac{1}{6}\ ^1N(01;01) -\sqrt{\frac{2}{9}}\ ^1N(01;21)+\frac{1}{3}\ ^1N(21;21)] \\ \frac{1}{16\pi}(1+\cos\vartheta)[2\ ^1N(01;01) +2\sqrt{2}\ ^1N(01;21)+\ ^1N(21;21)] \\ -\frac{1}{16\pi}(-1+\cos\vartheta)[2\ ^1N(01;01) +2\sqrt{2}\ ^1N(01;21)+\ ^1N(21;21)] \\ \frac{1}{16\pi}\sin\vartheta[-2\ ^1N(01;01) +\sqrt{2}\ ^1N(01;21)+2\ ^1N(21;21)] \end{array} \right )\ \ . \end{equation} The decomposition of the imaginary part of Eq.~(\ref{expans}) can now be written for the helicity state matrix-elements in matrix notation \begin{equation} {\bf N}^{\pi,\omega,A_1,H_1} =\left ( \begin{array}{ccccc} S_1 & V_1 & T_1 & P_1 & A_1 \\ S_2 & V_2 & T_2 & P_2 & A_2 \\ S_3 & V_3 & T_3 & P_3 & A_3 \\ S_4 & V_4 & T_4 & P_4 & A_4 \\ S_5 & V_5 & T_5 & P_5 & A_5 \end{array} \right ){\bf R}\equiv {\bf M} {\bf R} \end{equation} \noindent where ${\bf R}=(\rho_S \ \rho_V \ \rho_T \ \rho_P \ \rho_A)$. $S_i$, $V_i$, $T_i$, $P_i$, and $A_i$ are helicity state matrix-elements of the Fermi invariants using the same indexing as for the $N$ amplitudes and we find \begin{equation} {\bf M} =\left ( \begin{array}{ccccc} \beta^2-1 & -\cos\vartheta& -2\cos\vartheta & -\beta^2 & 1 \\ \beta^2-1 & -\cos\vartheta& 2\cos\vartheta(-2\beta^2+1) & \beta^2 & -1 \\ 0 & -\beta^2(1+\cos\vartheta) & -2(1+\cos\vartheta) & 0 & -(\beta^2-1)(1+\cos\vartheta) \\ 0 & -\beta^2(1-\cos\vartheta) & -2(1-\cos\vartheta) & 0 & (\beta^2-1)(1-\cos\vartheta) \\ 0 & \beta\sin\vartheta & 2\beta\sin\vartheta & 0 & 0 \end{array} \right ) \end{equation} The spectral functions are then simply obtained by calculating ${\bf R }={\bf M}^{-1}{\bf N}^{\pi,\omega,A_1,H_1}$ yielding the result of Eq.~(\ref{SpecFunc}). \end{appendix}
\section{Scaled Phase Space} The most important constraint on any calculation of pions from proton-antiproton annihilation at rest is that of energy momentum conservation. That is, the rate for finding $n$ pions of three-momenta $\mbox{\bf k}_i$, with $i=1...n$, should be proportional to the differential phase space factor $\rho_n(s, \{\mbox{\bf k}_i\})$ given by \begin{equation} \rho_n(s,\{\mbox{\bf k}_i\}) = \delta^4(k_t -\sum_{i=1}^n k_i) \prod_{i=2}^n \frac{d^3\mbox{\bf k}_i}{2 \omega_i} \end{equation} where $s = (k_t)^2$, $k_t$ is the total four-momentum of the annihilating pair and $\omega_i$ is the energy of the $i$-th pion, $\omega_i= \sqrt{\mbox{\bf k}_i^2+\mu^2}$, with $\mu$ the pion mass. For annihilation at rest $k_t = (2M, \mbox{\bf 0})$, with $M$ the nucleon mass. The ``phase space only" (PSO) assumption is that all other aspects of the annihilation process depend very weakly on $\{\mbox{\bf k}_i\}$ so that the entire dependence is given by $\rho_n(s,\{\mbox{\bf k}_i\})$. For example the single pion momentum spectrum for $n$ pions is obtained by integrating $\rho_n(s,\{\mbox{\bf k}_i\})$ over all but one of the final momenta. We will return below to the result of that integration. If we want to compare the branching ratio for $n$ to that for $n+1$ pions we observe that the corresponding integrals in (1) differ in dimension by two units of momentum. Thus to compare them, all other things being equal, we must construct a quantity of uniform dimension, $R(n)$, by scaling the total phase space for $n$ pions we write \begin{equation} R(n) = \frac{(L)^{2n}}{n!} \int \rho_n(s,\{\mbox{\bf k}_i\}) \end{equation} where $L$ is a length and an $n$ independent overall normalization factor is set equal to 1. We call this picture for relating multiplicities by a dimensional scaling ``scaled phase space," SPS. In terms of $R(n)$ one can calculate the average number of pions in the SPS picture by \begin{equation} \hat{n} = \frac{\sum_n n R(n)}{\sum_n R(n)} \end{equation} This $\hat{n}$ depends on $L$ which can be varied to give the experimental value, $\hat{n} = 5$. With the scaling length fixed by this empirical constraint, the resulting pion multiplicity distribution, the normalized $R(n)$, looks very much like the experimental one. Fig.~1 shows the pion multiplicity distribution calculated in the SPS formalism\footnote{The non-relativistic form of phase space leads to very similar distributions, as we have checked and as was already reported in \cite{Dover} } with $L = 1.2$ fm, adjusted to give the correct average pion number. It has a gaussian shape with an average of 5 (put in by hand) and a variance of 1. Note no statistical assumption has been made to obtain the observed gaussian distribution or the correct variance. Fig.~1 also displays the corresponding Skyrme-coherent state calculation of Sect.~2 and the measured distribution. Both calculations agree excellently with the data. \begin{figure}[htbp] \begin{center} \vspace*{-10mm} \mbox{\epsfysize=6cm\epsffile{cohst1.epsf}} \vspace*{5mm} \end{center} \caption{\label{Fig1}% Pion multiplicity distribution in proton antiproton annihilation at rest. The open dots are from the scaled phase space model and the solid dots the from Skyrme-coherent state approach. Both have been fixed to give an average pion multiplicity of 5. The histogram is the data presented in \protect\cite{Dover,Sedlak}. } \end{figure} In earlier treatments of the scaling model $L$ was connected to a volume \cite{Dover,old} by $V=(2 \pi L)^3$. The $2 \pi$ comes from the normalization of phase space density. The volume needed to give $\hat{n}=5$ turns out to be very large by nucleon standards, of order $(2\pi)^3$ fm$^3$. There is much discussion in the early literature about the meaning of this large volume \cite{Dover,old}. An alternate picture is that one should write $V = g^2 \cdot v$ where $v$ is a reasonable volume ($\sim \frac{4 \pi}{3}\mu^{-3}$ = $12.5$ fm$^3$) while $g$ is a dimensionless number that gives the amplitude for emitting one more pion. Now it is $g$ that is large. The notion that the amplitude for emitting $n$ pions should be proportional to $g^n$, the coupling constant to the $n$-th power, makes sense perturbatively, but is being employed here for $g$ large. Annihilation is nonperturbative. Thus the dynamical origin of the SPS picture remains obscure, even though it can certainly fit the pion multiplicity distribution with one free parameter. Historically statistical approaches were unable to account for the pion multiplicity distribution in a pions only picture with a ``reasonable" scaling parameter, or to connect the picture to more fundamental theory. Adding heavy mesons that decay sequentially to pions helps the phenomenology, but at the expense of more free parameters \cite{van}. There have also been attempts to make thermodynamic models of annihilation without imposing the constraints of energy-momentum conservation \cite{Blumel}. These do poorly for the pion multiplicity distribution even when $\hat{n}$ is fixed at 5. This serves to further emphasize the primary importance of the four-momentum phase space constraint. For fixed multiplicity, $n$, the PSO picture has no dynamics. Relating the amplitude for one $n$ to the next implies dynamical assumptions outside PSO as in the scaled phase space scheme SPS, discussed above. For a natural candidate that introduces genuine dynamics we turn to the Skyrme-coherent state approach. It relates the weights and momentum spectra of $n$ pion emission in a transparent way to the properties of the underlying annihilation process. \section{Skyrme-Coherent State Approach} In the Skyrme-coherent state approach one uses a classical meson field theory in which baryons appear as topologically stable solitons to model the dynamics of annihilation. This picture, invented by T.H.R. Skyrme \cite{Skyrme}, is connected to QCD in the limit of a large number of colors, and through that to the long wave length or nonperturbative limit of QCD \cite{tHooft,Witten}. It is found that in the Skyrme approach, annihilation proceeds very rapidly leading to a burst of relatively intense classical pion radiation \cite{SSLK,SWA}. To connect with the physical pion quanta of experiment, that classical wave is used to generate a quantum mechanical coherent state \cite{Klauder}. There are two steps to this process, the use of Skyrme dynamics to generate the classical pion radiation from annihilation, and the subsequent quantization of that radiation using coherent states. The circumstances of annihilation seem particularly well suited to this combination. A standard coherent state does not have fixed four-momentum, but as we saw above, that constraint is crucial. Hence the coherent state must be projected onto a state of definite energy-momentum \cite{HornSilver}, and if we are interested in pion charge ratios, a state of definite isospin \cite{UCSB}. We have developed the formalism for doing all this \cite{us}. A pion coherent state contains an exponential in the pion field creation operator. It is a single quantum state containing all pion numbers. This is the physics appeal of the coherent state approach, namely that all the pion channels are collected into a single state. Thus questions about the relation of the rate or spectrum in the $n$ pion channel to that in the $n+1$ channel, are naturally answered. No new parameters or assumptions are needed to address them. Since the coherent state approach can also be generalized to include energy-momentum conservation, the results discussed in Sect.~1 come out, but now with a clear origin for the relationship among the channels. In the Skyrme-coherent state approach, the difficult dynamics of nucleon-antinucleon interaction and subsequent annihilation into pions is done classically, and quantum mechanics only enters to describe the propagation of the coherent state after the fields have reached the radiation zone, where they are non-interacting. Although this program is far simpler than the corresponding full quantum QCD calculation of annihilation, it is still complicated to execute with the classical, nonlinear field equations of Skyrme and up to now has not been fully implemented. We have not studied the development of the annihilation process itself, but rather have begun with an assumed initial spherically symmetric configuration of classical pion field. It is in this initial pion configuration, that free parameters enter. The remaining dynamics is completely determined. We take a simple initial pion configuration characterized by a size and magnitude. The magnitude is fixed by the total energy of the system, $2M$, leaving only the size to be determined. This is fit to the average pion multiplicity, or equivalently, the inclusive single pion momentum distribution. We find a size of order 1~fm, a completely reasonable result. Note that if at some future date we are able to do the Skyrme calculation of annihilation from the beginning, there would be no free parameters whatsoever. The introduction of a finite source size for the pion radiation leads to a form factor for pion emission $f(k_i)$. This form factor is the Fourier transform of the classical asymptotic pion field. As such it is similar to, but not identical to the Fourier transform of the pion source distribution because of the intervening Skyrmion dynamics. For the coherent state formalism of \cite{us} we must replace (2) by \begin{equation} R^{COH}(n) = \frac{1}{n!} \int \prod_{i=2}^n |f(k_i)|^2 \rho_n(s,\{\mbox{\bf k}_i\}) \end{equation} (without the complication of isospin). This expression is completely specified by the coherent state formalism through the form factors. Hence in the Skyrme-coherent state approach the ratios of multiplicities are controlled not by some arbitrary volume, but by the intensity of the classical radiation field. In this work we take the analytic form for the form factor which we have used before \cite{us}. Adapted to the relativistic phase space convention it reads \begin{equation} |f(k)|^2 = \frac {2 C_0 \mbox{\bf k}^2}{ \omega (\mbox{\bf k}^2+\alpha^2)^2 (\omega^2+\alpha^2)^2} \label{fk} \end{equation} with $\alpha = 3.0 \mu$ and $C_0= 0.061$ GeV$^5$. $C_0$ has been determined by requiring that the energy release of the classical pion source is $2 M$. The size parameter $\alpha$ has been adjusted to the measured average multiplicity $\hat{n}=5$ imposing energy momentum-conservation\footnote{\label{a \mbox{Without energy momentum conservation the classical mean pion multiplicity} \mbox{$\int |f(k)|^2 d^3\mbox{\bf k} (2 \omega)^{-1}$} is $4.5$ for $\alpha = 3 \mu$.} as in \cite{us}. Note that any classical field theory capable of describing the evolution of classical pion radiation from annihilation will lead to a quantum coherent state and to form factors. We emphasize the Skyrme method because it is the only one we know that naturally gives annihilation and subsequent pion radiation. For the pion multiplicity spectrum we have seen in Fig.~1 that the introduction of form factors in the coherent state approach leads to results equivalent to those of the SPS model and in agreement with experiment. However now the relative weighting of different multiplicities has a simple physical origin in the strength and size of the initial pion source. The strength is determined by the energy release and the spatial distribution corresponding to (\ref{fk}) has an r.m.s. of 0.7~fm which is reasonable. Next we calculate the single pion momentum spectra for the pion multiplicities individually, by integrating the phase space over all but one of the pion momenta \begin{eqnarray} \frac{dN_n(K)}{dK} & = & \frac {1}{n!} \int \delta(K - |K_1|) \prod_{i=2}^n |f(k_i)|^2 \rho_n(s,\{\mbox{\bf k}_i\}) \\ \frac{dN(K)}{dK} & = & \sum_n \frac{dN_n(K)}{dK} \nonumber \end{eqnarray} where we use the notation $K = |\mbox{\bf k}| $. \begin{figure}[htbp] \begin{center} \vspace*{-12mm} \mbox{\epsfysize=12cm\epsffile{cohst2.epsf}} \vspace*{-18mm} \end{center} \caption{\label{Fig2}% The single pion inclusive momentum distribution ($dN_n/dK$) for annihilation to channels with pion multiplicity $n=3$ to $8$. The solid line is from the Skyrme-coherent state approach and the dotted line is phase space only. All distributions are normalized to one. } \end{figure} \begin{figure}[htbp] \begin{center} \mbox{\epsfysize=8cm\epsffile{cohst3.epsf}} \vspace*{-12mm} \end{center} \caption{\label{Fig3}% The single pion inclusive momentum distributions $dN_n/dK$ for each multiplicity weighted by the probability of that multiplicity and the sum of these, $dN/dK$, which is the full inclusive pion momentum spectrum. The Skyrme-coherent state case is shown on the left and the scaled phase space case on the right. The probabilities of the different multiplicities are shown on the graph. The summed spectrum is normalized to one. } \end{figure} \begin{figure}[htbp] \begin{center} \mbox{\epsfysize=8cm\epsffile{cohst4.epsf}} \vspace*{-12mm} \end{center} \caption{\label{Fig4}% The single pion momentum distribution, $dN/dK$, summed over all multiplicities, normalized to the total number of pions. The dotted line is from the scaled phase space model, the solid line from the Skyrme-coherent state approach and the data are from \protect\cite{Sedlak}. } \end{figure} In Fig.~2 we show the single particle spectra ($dN_n/dK$) for $n=3$ to $8$ final pions for the scaled phase space model SPS and for the coherent state approach \cite{us,LuAmado}. For each $n$, each graph is normalized to 1. We see that there is some difference in detail between the SPS and the coherent state approach, but they are quite similar in general shape. We have been unable to find recent data with which to compare these pion spectra for fixed $n$. In Fig.~3 we show the single pion momentum spectra of Fig.~2 with their correct relative weights. The probabilities of the different multiplicities, $P_n$ are listed on the figure. For the coherent state case the relative weights come out of the dynamics, for the phase space case they are put in through the scaling, fit to give $\hat{n} =5$. We see that in both cases only the $n=4,5,6$ multiplicities have substantial weight. Also shown in Fig.~3 are the weighted sum (normalized to one) of the momentum distributions in each multiplicity, the inclusive single pion momentum spectrum. In Fig.~4 we show that inclusive pion spectrum again (this time normalized to the total number of direct pions) comparing the SPS result, the coherent state result and the data reported in \cite{Dover}. The two calculations agree qualitatively, giving an equivalently good account of the data. \section{Conclusions} We have seen that the principal features of the pion multiplicity distribution and of the pion momentum spectrum in proton antiproton annihilation at rest can come from phase space so long as one connects the probabilities for different pion multiplicities. This can be done in an ad hoc way in the scaled phase space picture by introducing a scaling volume or in a dynamically motivated way in the context of the Skyrme-coherent state approach. The scaling volume needed in the phase space picture, $(2 \pi)^3$~fm$^3$, is an order of magnitude too large. No such volume interpretation is required in the Skyrme approach. Rather a form factor appears naturally the strength of which is fixed by the magnitude of the classical pion field, or equivalently by the energy released in annihilation, $2M$, and the range of which is fit to get an average pion number of 5, yielding a size of about $0.7$ fm. Furthermore if a complete calculation of annihilation using Skyrme dynamics were carried out (a difficult but not impossible task) there would be no free parameters in its description of annihilation. Finally we should point out that we have only discussed inclusive pion multiplicity and momentum spectra here. The Skyrme-coherent state picture has also been used to calculated pion charge branching ratios and extended to include vector mesons all with no new free parameters \cite{LuAmado}. These vector mesons are generated by the extended Skyrme dynamics since we take the initial configuration to be pions only. The calculated charge branching ratios and branching ratios into vector mesons ($\rho$ and $\omega$) are in qualitative agreement with experiment \cite{Dover,Sedlak}. A corresponding phase space only calculation would require additional free parameters to generate these branching ratios. Further afield, two pion correlations, which have been discussed in the Skyrme-coherent state picture \cite{ACDLL94,LA2}, find no natural explanation in the scaled phase space approach. \section{Acknowledgments} RDA, FC, and J-PD thank the theory group of the Division of Nuclear and Particle Physics of the Paul Scherrer Institute for, once again, providing a stimulating environment for this work. The work of RDA is partially supported by the United States National Science Foundation.
\section{INTRODUCTION} \label{sec:intro} It is roughly a decade since the $\xi(2220)$, now known as the $f_4(2220)$, was discovered by the MARK III collaboration in $J/\psi$ radiative decays to $K^+ K^-$ and $K_S K_S$ final states \cite{markiii}. Its most interesting property, which attracted considerable attention, was its narrow width of roughly 30~MeV. Because the width was inconsistent with expectations for a conventional $q\bar{q}$ meson with such a large mass, the $\xi$'s discovery led to speculation that it might be a Higgs boson \cite{higgs}, a bound state of coloured scalars \cite{scalar}, a four quark state \cite{4quark,pakvasa}, a $\Lambda \bar{\Lambda}$ bound state \cite{ono87}, a hybrid \cite{hybrid}, or a glueball \cite{glueball}. Despite the prevailing wisdom, the authors of Ref.~\cite{godfrey84,pakvasa} argued that the properties of the $\xi(2220)$ could be consistent with those of a conventional meson: the L=3 $s\bar{s}$ meson with $J^{PC}=2^{++}$ or $J^{PC}=4^{++}$. In the original analysis of L=3 $s\bar{s}$ properties it was shown that of the $q\bar{q}$ states with the appropriate $J^{PC}$ quantum numbers only the $^3F_2$ and $^3F_4$ $s\bar{s}$ states of the first L=3 multiplet have masses consistent with the $\xi(2220)$ \cite{godfrey84}. According to this analysis these two states were exceptional in that they have a limited number of available decay modes which are all relatively weak. However, the analysis was not exhaustive in that it did not calculate the decay widths to all possible final states. In particular it made the assumption, which we will see to be incorrect, that the decays to an $L=1$ meson and a $K$ or $\eta$ were small on the basis of phase space arguments alone. To further complicate the discussion, more recent experiments have observed a hadronic state decaying to $K\bar{K}$ in different reactions and with different properties. The various experimental results relevant to the $\xi(2220)$ are summarized in Table \ref{table1}. The most recent measurement of the $\xi(2220)$ properties by the BES collaboration \cite{bes} indicates that its decays are approximately flavour symmetric giving support to the glueball interpretation. At the same time, although the narrow $\xi(2220)$ was not seen in $J/\psi$ radiative decays by the DM2 experiment despite the fact that DM2 has slightly higher statistics, DM2 did observe a broader state decaying into $K\bar{K}$ \cite{dm2}. If all the experiments are taken at face value the overall picture is confused and contradictory. In this paper we re-examine the nature of the $\xi(2220)/f_4(2220)$ meson and calculate the partial widths of the $^3F_2$ and $^3F_4$ $s\bar{s}$ states to all OZI-allowed 2-body final states allowed by phase space. To give a measure of the reliability of our analysis we calculate the widths using both the $^3P_0$ decay model (often referred to as the quark-pair creation decay model) \cite{leyaouanc73,roberts} and the flux-tube breaking decay model \cite{kokoski87}. As an additional consistency check we calculated several partial widths using the pseudoscalar decay model \cite{godfrey85}. Our goal is to shed some light on the nature of the $\xi(2220)$ by comparing the quark model predictions for the hadronic widths to the various experimental results. The outline of the paper is as follows. In section \ref{sec2} we briefly outline the models of hadron decays and the fitting of the parameters of the models. We relegate the details to the appendices. In section \ref{sec3} we present the results of our calculations for the $L=3$ mesons and discuss our results. In the final section we attempt to make sense of the various contradictory experimental results and put forward our interpretation along with some suggested measurements which may clear up the situation. \section{MODELS OF MESON PROPERTIES AND DECAYS} \label{sec2} The quark model has proven to be a useful tool to describe the properties of hadrons. The quark model has successfully described weak, electromagnetic, and strong couplings \footnote{See for example Ref. \cite{godfrey85}.}. In some cases we will use simplified meson wavefunctions which have been used elsewhere to describe hadronic decays \cite{kokoski87} while in other cases we will use more complicated wavefunctions from a relativized quark model which includes one-gluon exchange and a linear confining potential \cite{godfrey85}. The strong decay analysis was performed using the QCD based flux-tube breaking model \cite{kokoski87}. It has the attractive feature of describing decay rates to all possible final states in terms of just one fitted parameter. We also include results for the $^3P_0$ model, often referred to as the quark-pair creation model \cite{leyaouanc73,roberts}, which is a limiting case of the flux-tube breaking model and which greatly simplifies the calculations and gives similar results. As a final check we calculated some partial widths using the pseudoscalar emission model \cite{godfrey85} and confirmed that it also gave results similar to those of the flux-tube breaking model. \subsection{Decays by the $^3P_0$ Model} The $^3P_0$ model \cite{leyaouanc73,roberts} is applicable to OZI-allowed strong decays of a meson into two other mesons, as well as the two-body strong decays of baryons and other hadrons. Meson decay occurs when a quark-antiquark pair is produced from the vacuum in a state suitable for quark rearrangement to occur, as in Fig.~\ref{3p0decay}. The created pair will have the quantum numbers of the vacuum, $^3P_0$. There is one undetermined parameter $\gamma$ in the model - it represents the probability that a quark-antiquark pair will be created from the vacuum. The rest of the model is just the description of the overlap of the initial meson (A) and the created pair with the two final mesons (B,C), to calculate the probability that rearrangement (and hence decay) will occur. A brief description of the model is included in Appendix \ref{appa}, and the techniques by which the calculations were performed are discussed in Appendices \ref{appc} and \ref{appd}. \subsection{Decays by the Flux-Tube Breaking Model} In the flux-tube picture a meson consists of a quark and antiquark connected by a tube of chromoelectric flux, which is treated as a vibrating string. For mesons the string is in its vibrational ground state. Vibrational excitations of the string would correspond to a type of meson hybrid, particles whose existence have not yet been confirmed. The flux-tube breaking decay model \cite{kokoski87} is similar to the $^3P_0$ model, but extends it by considering the actual dynamics of the flux-tubes. This is done by including a factor representing the overlap of the flux-tube of the initial meson with those of the two outgoing mesons. A brief review of the model is given in Appendix \ref{appb}, and the techniques by which the calculations were performed are discussed in Appendices \ref{appc} and \ref{appd}. \subsection{Fitting the Parameters of the Decay Models} The point of these calculations is to obtain a reliable estimate of the $^3F_2$ and $^3F_4$ $s\bar{s}$ meson decay widths. To do so we considered several variations of the flux-tube breaking model. By seeing how much the results vary under the various assumptions we can estimate the reliability of the predictions. The first variation lies with the normalization of the mock meson wavefunctions and the phase space used to calculate the decay widths \cite{geiger94}. In the Appendices we have normalized the mock meson wavefunctions relativistically to $2 E$ and used relativistic phase space, which leads to a factor of $E_B E_C/M_A$ in the final expression for the width in the centre of mass frame. We will refer to this as relativistic phase space/normalization (RPSN). However, there are arguments \cite{isgurpc} that heavy quark effective theory fixes the assumptions in the mock meson prescription and suggests that the energy factor be replaced by $\widetilde{M}_B \widetilde{M}_C/\widetilde{M}_A$, where the $\widetilde{M}_i$ are the calculated masses of the meson $i$ in a spin-independent quark-antiquark potential \cite{kokoski87}. (In other words $\widetilde{M}_i$ is given by the hyperfine averaged mass that is equal to the centre of gravity of the triplet and singlet masses of a multiplet of given $L$.) We will refer to this as the Kokoski-Isgur phase space/normalization (KIPSN). The second variation in our results is the choice of wavefunctions. We calculate decay widths for two cases. In the first we use simple harmonic oscillator (SHO) wavefunctions with a common oscillator parameter for all mesons. In the second case we use the wavefunctions, calculated in a relativized quark model, of Ref. \cite{godfrey85} which we will label RQM. In all we looked at six cases: the $^3P_0$ model using the SHO wavefunctions, the flux-tube breaking model again using the SHO wavefunctions, and the flux-tube breaking model using the RQM wavefunctions of Ref. \cite{godfrey85}; in all three cases we used both choices of phase space/normalization. Some comments about the details of the calculations are in order. For the SHO wavefunctions, we took for the oscillator parameter $\beta = 400$~MeV which is the value used by Kokoski and Isgur \cite{kokoski87}. However, different quark models find different values of $\beta$ so that there is the question of the sensitivity of our results to $\beta$. We will address this issue below. We used quark masses in the ratio $m_u:m_d:m_s = 3:3:5$ --- this differs from the calculations of Ref. \cite{kokoski87}, which ignored the strange-quark mass difference. In the RQM wavefunctions these parameters are already set --- the values of $\beta$ were found individually for each meson, and the quark masses were fitted: $m_u = 220$~MeV, $m_d = 220$~MeV, and $m_s = 419$~MeV. We have treated all mesons as narrow resonances, and have ignored mass differences between members of the same isospin multiplet \footnote{The one exception was for the decay $\phi \to K^+ K^-$ where the charged and neutral kaon mass difference is significant to the phase space.}. Masses were taken from the Review of Particle Properties 1994 \cite{pdb} if the state was included in their Meson Summary Table \footnote{The one exception was the $1^3P_0$ $s \bar{s}$ state --- see Table \ref{table4}.}. If it was not, then the masses predicted in Ref.~\cite{godfrey85} were used. (This includes the masses of the $1^3F_2$ and $1^3F_4$ $s\bar{s}$ mesons: 2240~MeV and 2200~MeV respectively.) Meson flavour wavefunctions were also taken from Ref.~\cite{godfrey85} - for the isoscalars we assumed ideal mixing ($\phi_{\rm non strange} = \frac{1}{\sqrt{2}}(u \bar{u}+d \bar{d})$, $\phi_{\rm strange} = s \bar{s}$), except for the radial ground state pseudoscalars, where we assumed perfect mixing ($\phi_\eta = \frac{1}{\sqrt{2}}(\phi_{\rm non strange}-\phi_{\rm strange})$, $\phi_{\eta'} = \frac{1}{\sqrt{2}}(\phi_{\rm non strange}+\phi_{\rm strange})$). We fitted $\gamma$, the one undetermined parameter of the model, in a global least squares fit of 28 of the best known meson decays. (We minimized the quantity defined by $\chi^2 =\sum_i (\Gamma^{model}_i - \Gamma^{exp}_i)^2/\sigma_{\Gamma_i}^2$ where $\sigma_{\Gamma_i}$ is the experimental error \footnote{For the calculations in the flux-tube breaking model, a 1\% error due to the numerical integration was added in quadrature with the experimental error.}.) The experimental values for these decays and the fitted values for the six cases are listed in Table \ref{table2}. To give a more descriptive picture of the results we plotted in Fig.~\ref{figure2}, on a logarithmic scale, the ratio of the fitted values to the experimental values. From Table \ref{table2} one can see that the results for the $^3P_0$ and flux-tube breaking models for the SHO wavefunctions are very similar\footnote {The one exception to this is the S-wave decay $K^*_0(1430) \to K\pi$ which seems particularly sensitive to the model.}. We therefore only plotted the $^3P_0$ model results using the SHO wavefunctions and the flux-tube breaking model results for the RQM wavefunctions. A reference line is drawn in each case for $\Gamma^{model}/\Gamma^{exp}=1$ to guide the eye. Since all the partial widths are proportional to $\gamma^2$, using a different fit strategy rescales $\gamma$. This is equivalent to simply shifting all points on the plot simultaneously making it easy to visualize any change in agreement for specific decays. The KIPSN gives a better overall fit to the data. Even so, certain decays, $K_3^*(1780) \to K\rho$ and $f_4(2050) \to \omega\omega$ for example, are fit much better using the RPSN. For both the RPSN and KIPSN one can see in Fig.~\ref{figure2} that a significant number of the decays differ from the experimental values by factors of two or more. Decays with two pseudoscalars in the final state tend to do better with the KIPSN but the KIPSN generally underestimates decays of high L mesons with vector mesons in the final states. On the other hand the RPSN tends to overestimate decays with two pseudoscalars in the final states. Similar observations can be made for the flux-tube breaking model using the RQM wavefunctions. Having said all this we stress that these are only general observations and exceptions can be found to any of them in Table \ref{table2}. One must therefore be very careful not to take the predictions at face value but should try if possible to compare the predicted decay to a similar one that is experimentally well known. Finally, we consider the sensitivity of our results to $\beta$. In addition to the fits discussed above, we performed simultaneous fits of both $\gamma$ and $\beta$ to the 28 decay widths for both the RPSN and the KIPSN. The resulting values of $\gamma$ and $\beta$ are 13.4 and 481~MeV respectively for RPSN and 5.60 and 371~MeV respectively for KIPSN. In both cases the overall fits improved slightly, with some widths in better agreement and some in worse agreement with experiment when compared to the fits for $\beta=400$~MeV. However, the fitted widths of the most relevant $^3F_4$ decays improve slightly for RPSN but show mixed results for KIPSN. We also redid our fits of $\gamma$ to the decay widths for $\beta=350$~GeV and $\beta=450$~GeV. For $\beta=350$~MeV the overall fit improves slightly for KIPSN although the predicted $f_4(2050)$ decay widths are a little worse and the $K_4(2045)$ widths are a little better. For RPSN the overall fit is a little worse as are the $^3F_4$ decays. For $\beta=450$~MeV the overall fit with KIPSN becomes a little worse as does the fitted $^3F_4$ widths while for RPSN the overall fit and fitted $^3F_4$ widths become a little better. We conclude that while there is some sensitivity to $\beta$, the results for modest changes in $\beta$ (including the $\beta$ we obtain by fitting $\gamma$ and $\beta$ simultaneously) are consistent with those for $\beta=400$~MeV within the overall uncertainty we assign to our results. It should be stressed that it is not sufficient to simply change $\beta$ but that a new value of $\gamma$ must be fitted to the experimental widths included in our fit. \section{RESULTS FOR $^3F_2$ and $^3F_4$ $\lowercase{s}\bar{\lowercase{s}}$ MESON DECAYS} \label{sec3} Using the $\gamma$'s obtained from our fit we calculated all kinematically allowed partial widths for the $^3F_2$ and $^3F_4$ $s\bar{s}$ meson decays. The results are given in Tables \ref{table3} and \ref{table4}. For the $^3F_4$ state the main decay modes are: \begin{equation} f_4' \to K^*(892) K^*(892), \; K\bar{K}, \; KK^*(892), \; \phi\phi, \; K K_2^*(1430), \; K K_1(1400), \; \eta\eta, \; \eta \eta' \end{equation} For the KIPSN and the SHO wavefunctions the total width is 132~MeV with the $^3P_0$ model. For this set of assumptions the $K\bar{K}$, $\eta\eta$, and $\eta\eta'$ modes are probably reasonably good estimates. However, the decay widths to $K K^*(892)$ and $K^*(892) K^*(892)$ are likely to be larger than the predictions. On this basis it does not seem likely to us that the $f_4'$ width is less than the predicted total width by a factor of two or more, i.e. we do not expect it to be less than about 70~MeV. If anything, we would expect it to be larger than the predicted width, i.e. $ > 140$~MeV. For the $^3F_2$ state we obtain results similar to the $^3F_4$ state for the $K\bar{K}$, $KK^*(892)$, and $K^*(892) K^*(892)$ modes. However, the $^3F_2$ also has large partial widths to $K K_1(1270)$, $K^*(892) K_1 (1270)$, $K K_2^*(1430)$ and $\eta f_1(1510)$. In fact, $K K_1(1270)$ is the dominant decay mode. It is large in all variations of the calculation we give in Table IV. The most closely related decay in our fit is the decay $\pi_2(1670) \to f_2(1270) \pi$ which is relatively large and is well reproduced by the KI normalization and SHO wavefunction case. The total width for this case is $\sim 400$~MeV \footnote{We note that the LASS collaboration has observed a $K_2^*(1980)$ state with a large total width of $373\pm 33 \pm 60$~MeV which could be associated with the strange meson partner of the $^3F_2 (s\bar{s})$ meson \cite{pdb}}. Even if this width is overestimated by a factor of two, it would still be too large to identify with the $\xi(2220)$. Although this result appears surprising it has a straightforward explanation. Examining Table \ref{table4}, the lowest angular momentum final states in $f_2'$ decay are P-waves. All of these decays are relatively broad but the $f_2' \to K_1(1270) K$ is the P-wave decay with the largest available phase space. In fact, one could almost order the P-wave decays using phase space alone. The analogous decay of the $f_4'$ is in an F-wave and therefore is subject to a larger angular momentum barrier. The lowest angular momentum partial wave for $f_4'$ decays is a D-wave which although it has the largest partial width of all $f_4'$ decays is still smaller than the P-wave $f_2'$ decay. As another measure of the reliability of these predictions we calculated the widths of the $K_4^*(2045)$ and $f_4(2050)$ mesons (the $^3F_4$ $K$-like and non-strange isovector mesons, respectively). The results for all significant kinematically allowed final states are given for the $^3P_0$ model using SHO wavefunctions in Tables \ref{table5} and \ref{table6} respectively. The results are consistent with the general fit results given in Table \ref{table2} and Fig.~\ref{figure2}. In general, the widths calculated using RPSN tend to be larger and those calculated using the KIPSN tend to be smaller. More specifically, decays to two pseudoscalar mesons using RPSN are generally overestimated while the results calculated using KIPSN are in reasonable agreement with experiment. There is no pattern for the decays to two vector final states. The decay $K_4^*(2045) \to K^*(892) \rho$ is greatly overestimated using RPSN but is in good agreement using KIPSN. In contrast, the predicted decay $f_4(2050) \to \omega\omega$ agrees well using RPSN but is greatly underestimated using KIPSN. The total widths tend to be overestimated using RPSN but are underestimated using KIPSN, both to varying degrees. The only conclusion we can draw from these results is that the total width probably lies between the two estimates but it is difficult to guess if it is closer to the lower or upper value. Finally, in Table VII we give the predicted total widths for the $^3F_2$ and $^3F_4$ $s\bar{s}$ states for the different values of $\beta$ considered in the previous section. Although they vary considerably, by roughly a factor of 2 going from $\beta=350$~MeV to $\beta=450$~GeV (except for the $\Gamma(^3F_2)$ with RPSN which varies by a factor of 3), these values are consistent within the large uncertainties we assign to our results. \section{DISCUSSION AND CONCLUSIONS} The motivation for this paper was to re-examine the possibility that the $\xi(2220)$ is an L=3 $s\bar{s}$ meson. This question is especially timely given the recent BES measurements of a narrow resonance with a mass of 2.2~GeV seen in $J/\psi$ radiative decays. To do so we calculated all kinematically allowed hadronic decays of the $^3F_2$ and $^3F_4$ $s\bar{s}$ states using several variations of the flux-tube breaking decay model. It appears very unlikely that the $\xi(2220)$ can be understood as the $^3F_2$ $s\bar{s}$ state. All variations of our calculation indicate that the $^3F_2$ $s\bar{s}$ is rather broad, $\gtrsim 400$~MeV. The dominant decay mode is the difficult to reconstruct $K K_1(1270)$ final state. Other final states with large branching ratios are $K^*(892) K_1(1270)$, $K K^*(892)$, $K^*(892) K^*(892)$, $K K_2^*(1430)$, $K\bar{K}$, and $\eta f_1(1510)$. It is more likely that the $^3F_4(s\bar{s})$ state can be associated with the $\xi(2220)$. The calculated width is $\sim 140$~MeV but given the uncertainties of the models it is possible, although perhaps unlikely, that the width could be small enough to be compatible with the width reported in the Review of Particle Properties 1994 \cite{pdb}. In this scenario the largest decay modes are to $K^*(892) K^*(892)$, $K \bar{K}$, $K K^*(892)$, and $\phi\phi$. Since only the $K\bar{K}$ final state has been observed an important test of this interpretation would be the observation of some of these other modes. There are, however, some problems with the $^3F_4(s\bar{s})$ identification of the $\xi(2220)$. Foremost is the flavour symmetric decay patterns recently measured by the BES collaboration \cite{bes}. These results contradict the expectations for a conventional $s\bar{s}$ meson. Second is the wide range of measured widths for this state. Although the Review of Particle Properties 1994 lists an average width of $38^{+15}_{-13}$~MeV the widths measured in hadron production experiments, LASS and E147, are larger while those measured in $J/\psi$ radiative decay tend to be narrow. The exception is the DM2 experiment which does not see, in $J/\psi$ radiative decay, a narrow state in $J/\psi \to \gamma K\bar{K}$ but does observe a relatively broad state at this mass. To account for these contradictions we propose a second explanation of what is being observed in this mass region --- that two different hadron states are observed, a narrow state produced in $J/\psi$ radiative decay and a broader state produced in hadron beam experiments. The broader state would be identified with the $^3F_4(s\bar{s})$ state. The predicted width is consistent with the quark model predictions and the LASS collaboration shows evidence that its quantum numbers are $J^{PC}=4^{++}$. We would then identify the narrow hadron state observed in the gluon rich $J/\psi$ radiative decays as a glueball candidate predicted by lattice gauge theory results \cite{lattice}. Recent lattice results indicate that glueballs may be narrower than one might naively expect \cite{lattice2}. The scalar glueball width is expected to be less that 200 MeV and one might expect a higher angular momentum state to be even narrower. The narrow state is not seen in hadron beam production because it is narrow, is produced weakly in these experiments through intermediate gluons, and is hidden by the $s\bar{s}$ state. Conversely, the broader state is not seen in $J/\psi$ radiative decays since this mode preferentially produces states with a high glue content. Crucial to this explanation is the experimental verification of the BES results on the flavour symmetric couplings of the state produced in $J/\psi$ radiative decay and the observation of other decay modes for the broader state in addition to the theoretical verification that the predicted tensor glueball is as narrow as the observed width. The $\xi(2220)$ has been a longstanding source of controversy. It is a dramatic reminder that there still is much that we don't understand about hadron spectroscopy and demonstrates the need for further experimental results to better understand this subject and ultimately better understand non-Abelian gauge theories, of which QCD is but one example. \acknowledgments This research was supported in part by the Natural Sciences and Engineering Research Council of Canada. S.G. thanks Nathan Isgur and Eric Swanson for helpful conversations.
\section*{Acknowledgements} D. G. B. was supported by CNPq and FUJB, Brasil, L. E. O. by CONICET , Argentina, and C. D. F. by ICTP, Italy. We acknowledge F. A. Schaposnik for useful comments. \newpage
\section{Introduction} Since the discovery of $p$-brane theories with manifest spacetime supersymmetry \cite{pol,berg}, it has become increasingly clear that there is a close relationship between such theories and the set of soliton-like solutions to supergravity theories \cite{town}. All the known supersymmetric $p$-brane theories achieve a matching of the on-shell world-volume bosonic and fermionic degrees of freedom, by virtue of a local fermionic symmetry known as $\kappa$ symmetry. This symmetry compensates for the excess of fermionic over bosonic degrees of freedom by gauging away half of the former. The consistency of $\kappa$ symmetry with spacetime supersymmetry places severe constraints on the spacetime dimension $D$ and the world volume dimension $d=p+1$ \cite{achu}. Four classic families of super $p$-branes were found to satisfy the consistency criterion. The members within each family are related by a process of double dimensional reduction \cite{dhis}, in which both the spacetime and the world volume are simultaneously compactified on a circle, and the dependence on the extra direction is dropped in each space. Thus the classic super $p$-branes may be classified by giving the maximal-dimensional member of each of the four families. These occur in $(D,d)= (11,3)$, $(10,6)$, $(6,4)$ and $(4,3)$. On a plot or `brane scan' of $D$ {\it vs} $d$, the additional $p$-branes obtained by double dimensional reduction lie on the North-east/South-west diagonal lines descending from the maximal cases. The idea that a super $p$-brane could be viewed as a long-wavelength description of a topological defect in a supersymmetric theory originated in the construction of the supermembrane in $D=4$ \cite{pol}. This supermembrane occurs as a kink solution of a $D=4$ chiral scalar supermultiplet theory with a potential giving a degenerate vacuum. A crucial feature of this solution is that half the original supersymmetry is left unbroken. This partial breaking of supersymmetry is also a general feature of all the subsequently-discovered $p$-brane solitons. Another feature of super $p$-branes became clear with the curved-superspace construction of the $D=11$ supermembrane action in \cite{berg}, and its generalisations to the other classic super $p$-branes. This new feature was the occurrence of integrability conditions on the supergravity background that are required for the existence of the world-volume $\kappa$ symmetry. In the case of the $D=11$ supermembrane, and of the type IIA string, related to it by double dimensional reduction, these integrability conditions imply the full set of supergravity field equations \cite{berg,dhis}. The association of super $p$-branes to supergravity is also natural because the supersymmetric $p$-branes can be viewed as the natural `matter' sources for the corresponding supergravity theories. A very specific r\^ole in this association is played by the antisymmetric tensor field strengths, whose gauge potentials couple directly to the $(p+1)$-dimensional world volumes. In the coupled solutions of super $p$-branes and their corresponding supergravity backgrounds, the backgrounds are naturally singular on the $p$-brane world volumes, which can act like delta-function sources. These singularities may or may not be clothed by horizons, depending upon the circumstances. Such singular supergravity solutions are called `elementary,' in distinction to the non-singular `solitonic' solutions described previously. The association of $p$-branes with singular supergravity solutions was made concrete with the explicit construction of superstring solutions in the case of $N=1$, $D=10$ supergravity \cite{dabl}. These solutions preserve half of the original supersymmetry, and consequently they saturate a Bogomol'ny bound on the energy density. Subsequently, an analogous elementary membrane solution of $D=11$ supergravity was found \cite{dust}. Many further solutions of supergravity theories have also been found, both for elementary $p$-branes \cite{pew1} and for solitonic $p$-branes \cite{pew2}. (There are also solitonic solutions in supergravity theories coupled to Yang-Mills, such as that based upon Yang-Mills instantons, and corresponding to the heterotic string \cite{pew3}.) The multiplicity of elementary and solitonic $p$-brane solutions to supergravity theories, covering many more values of $(D,d)$ than the classic $\kappa$-symmetric points on the brane scan, suggests that the original classification needs to be generalised. Leaving aside for the moment the problem of formulating more general $\kappa$-symmetric actions, it is worthwhile to try to find the general pattern of elementary and solitonic $p$-brane solutions in supergravity theories. Many supergravity theories in $D\le10$ dimensions can be obtained from $D=11$ supergravity by Kaluza-Klein dimensional reduction, in which a consistent truncation of the higher-dimensional to the lower-dimensional theory is made. Since the truncation is consistent, it follows that solutions of the lower-dimensional theory are also solutions of the higher-dimensional one. This lifting of solutions to the higher dimension is known as dimensional oxidation. In some cases, an elementary or solitonic brane solution in the lower dimension oxidizes to another elementary or solitonic brane solution in the higher dimension. The ability to view an oxidized brane solution as itself being a brane solution depends upon whether the isotropicity of the lower-dimensional solution extends to an isotropicity in the higher-dimensional sense. For the isotropicity to extend, the extra coordinate of the higher-dimensional spacetime must either become isotropically grouped with the $p$-brane coordinates of the lower dimension, making a $(p+1)$-brane, or else it must become isotropically grouped with the coordinates of the transverse space, making a $p$-brane in the higher dimension. As we shall show later, the latter can never happen within the framework of Kaluza-Klein dimensional reduction. The former, on the other hand, can occur under certain circumstances. This is the direct analogue, at the level of solutions to supergravity theories, of the process of double dimensional reduction of $p$-brane actions \cite{dhis}. Just as for those actions, it is useful in classifying the brane solutions to distinguish between the ones that can be oxidized to isotropic brane solutions of a higher-dimensional supergravity theory, and those that cannot be isotropically oxidized. We shall call the former solutions `rusty,' and the latter solutions `stainless.' Thus when constructing a brane scan of supergravity solutions, one may omit the rusty solutions, which are simply the Kaluza-Klein descendants of stainless solutions in some higher dimension. The full solution set is thus characterised by the stainless solutions. A frequently-encountered contention in the recent literature is that the only fundamental brane solutions occur in $D=11$ and $D=10$ supergravities, and that all the others are simply obtained by dimensional reduction. In this paper, we shall show that this is not the case, given our requirement of isotropicity in the oxidation process.\footnote{ An opposite viewpoint is to regard all oxidations of brane solutions as branes in the higher dimension. We prefer not to adopt this viewpoint since, if the isotropicity requirement on the world volume is dropped, the solutions are not ordinary extended objects, and moreover it would not then be clear what degree of anisotropicity should be regarded as acceptable.} In particular, we shall find new stainless brane solutions to supergravity theories in all $5\le D\le9$. (We shall not be concerned in the present paper with supersymmetric $p$-brane solutions to super Yang-Mills or other rigid supersymmetric theories.) Amongst other stainless examples, we shall find a 6-brane and a 5-brane in $D=9$, and a string in $D=5$, none of which are obtainable from $D=11$ or $D=10$ $p$-brane solutions by dimensional reduction. \section{Solutions and Kaluza-Klein dimensional reduction} \subsection{$p$-brane solutions} We are concerned with elementary and solitonic solutions of supergravity theories that admit interpretations as $p$-branes embedded in spacetime. These solutions will in general involve the metric tensor $g_{\sst{MN}}$, a dilaton $\phi$ and an $n$-index antisymmetric tensor $F_{\sst{M_1\cdots M_n}}$ in $D$ dimensions. The Lagrangian for these fields takes the form \begin{equation} {\cal L} = e R - \ft12 e(\partial\phi)^2 - {1\over 2\, n!} e e^{-a \phi} F^2\ , \label{boslag} \end{equation} where $e=\sqrt{-g}$ is the determinant of the vielbein. The equations of motion are \begin{eqnarray} \mathord{\dalemb{6.8}{7}\hbox{\hskip1pt}} \phi &=& -{a\over 2 n!}\, e^{-a \phi} F^2\ ,\nonumber\\ R_{\sst{MN}} &=& \ft12 \partial_{\sst{M}}\phi\,\partial_{\sst{N}}\phi + S_{\sst{MN}} \ ,\label{eqmo1}\\ \partial_{\sst{M_1}} (e e^{-a \phi} F^{\sst{M_1\cdots M_n}}) &=& 0\ ,\nonumber \end{eqnarray} where $S_{\sst{MN}}$ is a symmetric tensor given by \begin{equation} S_{\sst{MN}} = {1\over 2 (n-1)!}\, e^{-a \phi}\,\Big( F^2_{\sst{MN}} - \fft{n-1}{n(D-2)} F^2 g_{\sst{MN}}\Big) \ .\label{smndef} \end{equation} The ansatz for the metric for the $D$ dimensional spacetime is given by \cite{dabl,dkl} \begin{equation} ds^2 = e^{2A}\, dx^\mu dx^\nu \eta_{\mu\nu} + e^{2B}\, dy^m dy^n \delta_{mn}\ ,\label{metrform} \end{equation} where $x^{\mu}$ $(\mu = 0, \ldots, d-1)$ are the coordinates of the $(d-1)$-brane world volume, and $y^m$ are the coordinates of the $(D-d)$-dimensional transverse space. The functions $A$ and $B$ depend only on $r=\sqrt{y^my^m}$. Note that the form of the metric ansatz is preserved under the replacement $r\longrightarrow 1/r$. The Ricci tensor for the metric (\ref{metrform}) is given by \begin{eqnarray} R_{\mu\nu} &=& -\eta_{\mu\nu} e^{2(A-B)} \Big ( A'' + d {A'}^2 + {\tilde d} A' B' + \fft{{\tilde d}+1}{r}\, A'\Big)\ ,\nonumber\\ R_{mn} &=& -\delta_{mn} \Big(B'' + d A' B' + {\tilde d} {B'}^2 + \fft{2{\tilde d} +1}r\, B' + \fft{d}r\, A'\Big) \label{ricci}\\ && -{y^my^n \over r^2} \Big({\tilde d} B'' + d A'' - 2d A'B' + d {A'}^2 - {\tilde d} {B'}^2 -\fft{{\tilde d}}{r}\, B' - \fft{d}{r} \, A'\Big)\ ,\nonumber \end{eqnarray} where ${\tilde d}= D -d - 2$ and a prime denotes a derivative with respect to $r$. A convenient choice of vielbein basis for the metric (\ref{metrform}) is $e^{\underline \mu} = e^{A} d x^\mu$ and $ e^{\underline m} = e^{B} dy^m$, where underlined indices denote tangent space components. The corresponding spin connection is \begin{eqnarray} \omega^{{\underline\mu}{\underline n}} &=& e^{-B}\partial_n A\, e^{\underline \mu} \ ,\qquad \omega^{{\underline \mu}{\underline \nu}} = 0\ ,\nonumber\\ \omega^{{\underline m}{\underline n}} &=& e^{-B} \partial_n B\, e^{\underline m} -e^{-B} \partial_m B\, e^{\underline n}\ .\label{spincon0} \end{eqnarray} For the elementary $p$-brane solutions, the ansatz for the antisymmetric tensor is given in terms of its potential, and takes the form \cite{dabl} \begin{equation} A_{\mu_1\ldots\mu_{n-1}} = \epsilon_{\mu_1\ldots\mu_{n-1}} e^C \ ,\label{eleans} \end{equation} and hence \begin{equation} F_{m\mu_1\ldots\mu_{n-1}} = \epsilon_{\mu_1\ldots\mu_{n-1}} \partial_m e^C\ , \label{eleans2} \end{equation} where $C$ is a function of $r$ only. Here and throughout this paper $\epsilon_{\sst{M\cdots N}}$ and $\epsilon^{\sst{M\cdots N}}$ are taken to be the tensor densities of weights $-1$ and 1 respectively, with purely numerical components $\pm 1$ or $0$. Note in particular that they are not related just by raising and lowering indices using the metric tensor. The dimension of the world volume is given by $d=n-1$ for the elementary $p$-brane solutions. For the solitonic $(d-1)$-brane solutions, the ansatz for the antisymmetric tensor is \cite{pew2} \begin{equation} F_{m_1\cdots m_n} = \lambda \epsilon_{m_1\cdots m_n p }\, {y^p\over r^{n+1}} \ ,\label{solans} \end{equation} where $\lambda$ is a constant. The power of $r$ is governed by the requirement that $F$ should satisfy the Bianchi identity. The dimension of the world volume is given by $d=D-n-1$ for the solitonic $p$-brane solutions. For both types of solution, the symmetric tensor $S_{\sst{MN}}$ takes the form \begin{eqnarray} S_{\mu\nu} &=& -\fft{{\tilde d}}{2(D-2)}\, S^2\, e^{2(A-B)} \eta_{\mu\nu} \ ,\nonumber\\ S_{mn} &=& \fft{d}{2(D-2)}\, S^2 \,\delta_{mn} - \ft12\, S^2\, {y^my^n \over r^2} \ .\label{smnform} \end{eqnarray} The function $S$ is given in the two cases by \begin{eqnarray} {\rm elementary:}&& S = e^{-\ft12 a\phi - d A + C}\, C'\qquad \ \ \ d=n-1\ , \nonumber\\ {\rm solitonic:}&& S = \lambda e^{-\ft12 a\phi - {\tilde d} B}\, r^{-{\tilde d}-1} \qquad d= D-n-1\ .\label{sforms} \end{eqnarray} With these ans\"atze, the equations of motion for the dilaton and the metric tensor in (\ref{eqmo1}) become \begin{eqnarray} \phi'' + d A' \phi' + {\tilde d} B' \phi' + \fft{{\tilde d} +1}r\, \phi' &=&\ft12 \epsilon a S^2\ ,\nonumber\\ A'' + d {A'}^2 + {\tilde d} A' B' + \fft{{\tilde d}+1}r \, A' &=& \fft{{\tilde d}}{2(D-2)} S^2\ ,\nonumber\\ B'' + d A' B' + {\tilde d} {B'}^2 + \fft{2{\tilde d} +1}r \, B' + \fft{d}r\, A' &=& -\fft{d}{2(D-2)} S^2 \ ,\label{eqmo2}\\ {\tilde d} B'' + d A'' - 2 d A' B' + d {A'}^2 - {\tilde d} {B'}^2 - \fft{{\tilde d}}r \, B' - \fft{d}r \, A' + \ft12 {\phi'}^2 &=& \ft12 S^2\ ,\nonumber \end{eqnarray} where $\epsilon = 1$ for the elementary ansatz and $\epsilon = -1$ for the solitonic ansatz. The equation of motion for the field strength $F$ in (\ref{eqmo1}) is automatically satisfied by the solitonic ansatz (\ref{solans}), whilst for the elementary ansatz (\ref{eleans}) it gives rise to the equation \begin{equation} C'' + C'(C' + {\tilde d} B' - dA' - a \phi') + \fft{{\tilde d}+1}r \, C' = 0\ . \label{ceq} \end{equation} Solutions to the equations of motion (\ref{eqmo2}) and (\ref{ceq}) can be obtained by making the following ansatz: \begin{equation} A' =\fft{\epsilon}{\Lambda}\, S\ ,\qquad \phi' = \fft{\epsilon(D-2)a}{{\tilde d}} \, A'\ ,\label{soluans} \end{equation} where $\Lambda$ is a constant. By choosing $\Lambda$ such that \begin{equation} \Lambda^2 = \fft{(D-2)^2 a^2}{{\tilde d}^2} + \fft{2d(D-2)}{{\tilde d}}\ ,\label{betaeq} \end{equation} one can eliminate the non-linear terms ${A'}^2$, ${B'}^2$ and $A' B'$ from a linear combination of the last three equations in (\ref{eqmo2}).\footnote{There are more general solutions of the equations (\ref{eqmo2}) than those that follow from the ansatz (\ref{soluans},\ref{betaeq}). However, as we shall see later, when one considers supergravity theories the equations implied by requiring that half the superymmetry be preserved are equivalent to (\ref{soluans},\ref{betaeq}).} Then it is a simple matter to solve the equations; the solution is given by \begin{eqnarray} B&=&-\fft{d}{{\tilde d}} \, A \ , \qquad \phi = \fft{a(D-2)}{\epsilon {\tilde d}} \, A \ ,\nonumber\\ e^{-c A} &=& 1 + \fft{k}{r^{{\tilde d}}}\ ,\label{solution1} \end{eqnarray} where $k = \epsilon \Lambda\lambda/(2(D-2))$ and $ c= d + a^2 (D-2)/(2{\tilde d})$. In the elementary case, the function $C$ satisfies the equation \begin{equation} (e^C)' = \lambda\, e^{2cA}\, r^{-{\tilde d} -1}\ .\label{csol} \end{equation} In presenting these solutions we have chosen simple values for some integration constants where no loss of generality is involved. The solutions (\ref{solution1}) are valid when $d{\tilde d}>0$. For the cases $d=0$ or ${\tilde d}=0$, the solutions can also be straightforwardly obtained; an example will be given in section 4.2. Note that the forms of the metrics for both elementary and solitonic $(d-1)$-branes are the same, although, as we saw earlier, the solutions are obtained from a $(d+1)$-form antisymmetric tensor field strength in the former case, and from a $(D-d-1)$-form antisymmetric tensor field strength in the latter case. So far, we have obtained solutions for the bosonic theory described by the Lagrangian (\ref{boslag}) for arbitrary values of the constant $a$, and with an antisymmetric tensor of arbitrary degree. In supergravity theories, however, there occur antisymmetric tensors of certain specific degrees only, each with its corresponding specific value of the constant $a$. We may summarise the $a$ values arising in supergravity theories as follows. Without loss of generality, we may discuss all theories in versions where all antisymmetric tensor field strengths have degrees $n\le D/2$. The $a$ values are given by \begin{equation} a^2 = \Delta -\fft{2d{\tilde d}}{D-2}\ ,\label{avalue} \end{equation} where \begin{equation} d{\tilde d}=(n-1)(D-n-1)\ .\label{ndep} \end{equation} Some examples of values of $\Delta$ that arise in supergravity theories are $\Delta =4$ for $n \ne 2$, and $\Delta = 4$ and 2 for $n=2$. (See \cite{SS}, where a large class of supergravity theories in various dimensions can be found.) We shall discuss the set of $\Delta$ values in more detail in section 4.1. Note that in cases where there is no dilaton, the solution for the $A$ and $B$ functions that appear in the metric ansatz is precisely given by (\ref{solution1}) with the value of $a$ taken to be zero. In this sense we can assign the value $\Delta = 2d{\tilde d}/(D-2)$, which, by eqn (\ref{avalue}), sets $a=0$, in a supergravity theory where there is no dilaton. For example $\Delta=4$ for the 4-form field strength in $D=11$ supergravity, $\Delta=2$ for the 3-form field strength in $D=6$ self-dual supergravity, and $\Delta=\ft43$ for the 2-form field strength in $D=5$ simple supergravity. It follows from eqn (\ref{solution1}) that the metrics for the brane solutions are given by \begin{equation} ds^2 = \Big(1+\fft{k}{r^{\tilde d}}\Big)^{-\ft{4{\tilde d}}{(D-2)\Delta}}\, dx^\mu dx^\nu \eta_{\mu\nu} + \Big(1+\fft{k}{r^{\tilde d}}\Big)^{\ft{4d}{(D-2)\Delta}} \, dy^m dy^m\ .\label{metrsol} \end{equation} This coincides with the results given in ref.\ \cite{dkl} for the case of $\Delta=4$. Note from (\ref{betaeq}) and (\ref{solution1}) that in terms of $\Delta$, the functions $A$, $B$ and $\phi$ satisfy \begin{equation} A'=\fft{\epsilon{\tilde d}}{(D-2)\sqrt\Delta} \, S\ ,\qquad B'=- \fft{\epsilon d}{(D-2)\sqrt\Delta} \, S\ , \qquad\phi'=\fft{a}{\sqrt\Delta} \, S\ , \label{abphirel} \end{equation} and the dilaton is given by $e^\phi=(1+ kr^{-\tilde d})^{-2a/\Delta \epsilon}$ with $k=\ft12 \sqrt{\Delta}\lambda/\tilde d$. As we shall see in detail in the next section, some of the $(d-1)$-brane solutions that we have obtained in a $D$-dimensional supergravity can be isotropically oxidized to $d$-brane solutions in a $(D+1)$-dimensional supergravity. The degree of the antisymmetric tensor involved in a $p$-brane solution, and the value of the constant $a$, play crucial r\^oles in determining whether the solution can or cannot be isotropically oxidized in this way. At this point, a remark about supersymmetry is in order. In order for the solutions that we have obtained above to acquire an interpretation as {\it super} $(d-1)$-branes embedded in $D$-dimensional spacetime, we shall have to verify that these solutions preserve half of the supersymmetry of the corresponding supergravity theories. We have verified, case by case, that this is indeed true, at least as long as the antisymmetric tensor is part of the supergravity multiplet. In fact, the conditions arising from the requirement of preserving half of the supersymmetries turn out to be precisely equivalent to those that we imposed in the ansatz (\ref{soluans}). In concluding this subsection, we return to a more detailed discussion of a point to which we alluded earlier, namely that we may choose, when discussing the solution set of elementary and solitonic branes in supergravity theories, to restrict our attention to the versions of the various supergravity theories in which all antisymmetric tensors $F_n$ have degrees $n$ that do not exceed $D/2$. The reason why we may do this without losing generality is that an elementary or solitonic solution of a version of a supergravity theory in which the antisymmetric tensor participating in the solution is dualised is {\it precisely} the same as the solitonic or elementary solution, respectively, of the undualised form of the supergravity theory. To see this, consider the solitonic solution of (\ref{eqmo1}), with $F_n$ given by the ansatz (\ref{solans}). This has \begin{equation} F_n=\fft{1}{n!} F_{m_1\cdots m_n}\,dy^{m_1}\wedge\cdots\wedge d y^{m_n}= \fft{\lambda}{n!} e^{-nB}\, \epsilon_{m_1\cdots m_n p}\, \fft{y^p}{r^{n+1}} \, e^{{\underline m}_1}\wedge \cdots \wedge e^{{\underline m_n}}\ . \end{equation} Thus the Hodge dual of this $n$-form is given by \begin{equation} *F_n= \fft{\lambda}{(D-n)!}\, \fft{y^m}{r^{n+1}}\, e^{-nB}\, \epsilon_{\mu_1\cdots \mu_d} \, e^{\underline m}\wedge e^{{\underline\mu}_1}\wedge \cdots \wedge e^{{\underline \mu}_d}\ .\label{fdual} \end{equation} In the dual version of the theory, the $(D-n)$-form $\widetilde F$ whose Bianchi identity implies the field equation for $F_n$ given in (\ref{eqmo1}) is $\widetilde F= e^{-a\phi} *F_n$, which, from (\ref{fdual}), has components given by \begin{equation} {\widetilde F}_{m\mu_1\cdots \mu_d} = \fft{\lambda y^m}{r^{n+1}} \, e^{dA-{\tilde d} B -a\phi}\, \epsilon_{\mu_1\cdots \mu_d}\ . \end{equation} Hence by using (\ref{solution1}), with ${\tilde d}=n-1$, we see that ${\widetilde F}_{m\mu_1\cdots \mu_d}$ is precisely of the form of the elementary ansatz (\ref{eleans}) for a $(d+1)$-index field strength, where the function $C$ satisfies its equation of motion (\ref{csol}). Thus we see that the solitonic solution of the dualised theory is precisely the same thing as the elementary solution of the undualised theory, and {\it vice versa}, with the antisymmetric tensor written in different variables. We may therefore, without loss of generality, consider all supergravity theories in their versions where the degrees of their antisymmetric tensors $F_n$ satisfy $n\le D/2$. The set of all elementary and solitonic brane solutions of these theories spans the entire set of inequivalent brane solutions of these theories together with their dualised versions. \subsection{Kaluza-Klein dimensional reduction} In order to describe the processes of oxidation and reduction, we need to set up the Kaluza-Klein procedure for dimensional reduction from $(D+1)$ to $D$ dimensions. Let us denote the coordinates of a $(D+1)$-dimensional spacetime by $x^{\hat\sst{M}}=(x^{\sst{M}}, z)$, where $z$ is the coordinate of the extra dimension. The $(D+1)$-dimensional metric $d\hat s^2$ is related to the $D$-dimensional metric $ds^2$ by \begin{equation} d\hat s^2 = e^{2\alpha\varphi} ds^2 + e^{2\beta\varphi} (dz + {\cal A}_{\sst{M}} dx^{\sst{M}})^2\ ,\label{kkans} \end{equation} where $\varphi$ and ${\cal A}$ are taken to be independent of the extra coordinate $z$. The constants $\alpha$ and $\beta$ will be determined shortly. A convenient choice for the vielbein $\hat e^{\hat\sst{A}}{}_{\hat\sst{M}}$ of the $(D+1)$-dimensional spacetime is \begin{eqnarray} \hat e^{\sst{A}}{}_{\sst{M}} = e^{\alpha\varphi} \, e^{\sst{A}}{}_{\sst{M}} \ ,&& \hat e^{\underline z}{}_{\sst{M}} = e^{\beta\varphi}\, {\cal A}_{\sst{M}} \ , \nonumber\\ \hat e^{\sst{A}}{}_{z} = 0\ , && \hat e^{\underline z}{}_z = e^{\beta\varphi}\ {}. \end{eqnarray} Note that $M$ and $z$ denote world indices, whilst $A$ and $\underline z$ denote tangent-space indices. The spin connection is given by \begin{eqnarray} \hat \omega^{\sst{AB}} &=& \omega^{\sst{AB}} + \alpha e^{-\alpha\varphi}\, \Big(\partial^{\sst{B}}\varphi\, \hat e^{\sst{A}} - \partial^{\sst{A}}\varphi\, \hat e^{\sst{B}}\Big) -\ft12 {\cal F}^{\sst{AB}} e^{(\beta-2\alpha)\varphi} \, \hat e^{\underline z}\ ,\nonumber\\ \hat\omega^{{\sst{A}}\underline z} &=& -\beta e^{-\alpha\varphi}\, \partial^{\sst{A}} \varphi\, \hat e^{\underline z} -\ft12 {\cal F}^{\sst{A}}{}_{\sst{B}} \, e^{(\beta-2\alpha)\varphi}\, \hat e^{\sst{B}}\ ,\label{spincon} \end{eqnarray} where $\partial_{\sst{A}}= E_{\sst{A}}{}^{\sst{M}} \partial_{\sst{M}}$ is the partial derivative with a tangent-space index, and ${\cal F}_{\sst{MN}}= 2\partial_{[\sst{M}} {\cal A}_{\sst{N}]}$. Here, $E_{\sst{A}}{}^{\sst{M}}$ is the inverse vielbein in $D$ dimensions. Choosing $\beta = -(D-2) \alpha $, we find that the $(D+1)$-dimensional Einstein-Hilbert action $\hat e\, \hat R$ reduces to \begin{equation} \hat e \hat R = e R -(D-1)(D-2) \alpha^2\, e (\partial\varphi)^2 - \ft14 e\, e^{-2(D-1)\alpha\varphi}\, {\cal F}^2\ .\label{ricscal} \end{equation} The Kaluza-Klein dilaton $\varphi$ may be given its canonical normalisation by choosing the constant $\alpha$ such that \begin{equation} \alpha^2= \fft{1}{2(D-1)(D-2)}\ .\label{alphaval} \end{equation} It is sometimes useful to have expressions for the $(D+1)$-dimensional Ricci tensor. Its tangent-space components are given, after setting $\beta=-(D-2)\alpha$, by \begin{eqnarray} \hat R_{\sst{AB}} &=& e^{-2\alpha\varphi} \Big(R_{\sst{AB}} - (D-1)(D-2)\alpha^2\, \partial_{\sst{A}}\varphi\, \partial_{\sst{B}}\varphi - \alpha\, \mathord{\dalemb{6.8}{7}\hbox{\hskip1pt}}\varphi \, \eta_{\sst{AB}}\Big) - \ft12 e^{-2D\alpha\varphi}\, {\cal F}_{\sst{A}}{}^{\sst{C}} {\cal F}_{\sst{ BC}}\ ,\nonumber\\ {\hat R}_{{\sst{A}}\underline z} &=& \ft12 e^{(D-3)\alpha\varphi}\, \nabla^{\sst{B}}\Big( e^{-2(D-1)\alpha\varphi}\, {\cal F}_{\sst{AB}} \Big)\ , \label{ricten}\\ \hat R_{\underline z \underline z} &=& (D-2)\, \alpha\, e^{-2\alpha\varphi}\, \mathord{\dalemb{6.8}{7}\hbox{\hskip1pt}} \varphi +\ft14 e^{-2D\alpha\varphi}\, {\cal F}^2\ .\nonumber \end{eqnarray} Let us now apply the above formalism to the case of a bosonic Lagrangian of the form (\ref{boslag}), but in $(D+1)$ rather than $D$ dimensions: \begin{equation} {\cal L} = \hat e \hat R -\ft12 \hat e (\partial \hat\phi)^2 - \fft1{2\, n!} \hat e e^{-\hat a\hat\phi}\, {\hat F_n}^2\ ,\label{higherbos} \end{equation} where we add a subscript index $n$ to indicate that $F$ is an $n$-form. The Kaluza-Klein ansatz for $\hat\phi$ is simply $\hat\phi=\phi$, where $\phi$ is independent of the extra coordinate $z$. For $\hat F_n$, which is written locally in terms of a potential $\hat A_{n-1}$ as $\hat F_n=d \hat A_{n-1}$, the ansatz for $\hat A_{n-1}$ is \begin{equation} \hat A_{n-1}=B_{n-1} + B_{n-2}\wedge dz\ ,\label{aans} \end{equation} where $B_{n-1}$ and $B_{n-2}$ are potentials for the $n$-form field strength $G_n=dB_{n-1}$ and the $(n-1)$-form field strength $G_{n-1}=d B_{n-2}$ in $D$ dimensions. Defining \begin{equation} G'_n=G_n - G_{n-1}\wedge {\cal A}\ , \end{equation} where ${\cal A}={\cal A}_{\sst{M}} dx^{\sst{M}}$, one finds \begin{equation} \hat F_n = {G'}_n + G_{n-1}\wedge (dz+ {\cal A})\ .\label{transgress} \end{equation} The tangent-space components of $\hat F_n$ in $(D+1)$ dimensions are therefore given by $\hat F_{\sst{A_1\cdots A_n}}= {G'}_{\sst{A_1\cdots A_n}} e^{-n\alpha\varphi}$ and $\hat F_{{\sst{A_1\cdots A_{n-1}}} {\underline z}}= G_{\sst{A_1\cdots A_{n-1}}} e^{-(n-1)\alpha\varphi -\beta\varphi}$. Substituting into (\ref{higherbos}), and using $\beta=-(D-2)\alpha$, we obtain the reduced $D$-dimensional Lagrangian \begin{eqnarray} {\cal L} &=& e R -\ft12 e (\partial\phi)^2 -\ft12 e (\partial\varphi)^2 - \ft14 e e^{-2(D-1)\alpha\varphi} \, {\cal F}^2 \nonumber\\ &&-\fft{e}{2\, n!} e^{-2(n-1)\alpha\varphi-\hat a \phi}\, {G'}_n^2 - \fft{e}{2\, (n-1)!} e^{2(D-n)\alpha\varphi -\hat a\phi}\, G_{n-1}^2 \ , \label{bosred} \end{eqnarray} where $\alpha$ is given by (\ref{alphaval}). As one sees, different combinations of $\varphi$ and $\phi$ appear in the exponential prefactors of ${G'}_n^2$ and $G_{n-1}^2$. Nonetheless, each of these prefactors may easily be seen to be of the form $e^{-a_n\tilde\phi}$, where $\tilde\phi$ is an $SO(2)$ rotated combination of $\varphi$ and $\phi$. In these prefactors, the coefficients $a_n$ satisfy the formula (\ref{avalue}) in $D$ dimensions, with $d{\tilde d}$ given by (\ref{ndep}), and with the {\it same} value of $\Delta$ as for $\hat a$ in $(D+1)$ dimensions. (Note that $d{\tilde d}$ in (\ref{avalue}) is $n$-dependent, so one obtains different values for the ${G'}_n^2$ and $G_{n-1}^2$ prefactors.) The 2-form field strength ${\cal F}$ has an $a$ value given by (\ref{avalue}) with $\Delta=4$. Most supergravity theories can be obtained from 11-dimensional supergravity {\it via} Kaluza-Klein dimensional reduction. Any such dimensional reduction can be viewed as a sequence of reductions by one dimension at a time, of the kind we are discussing here. Any solution of a lower-dimensional supergravity theory in such a sequence can therefore be reinterpreted as a solution of any one of the higher theories in the sequence by use of the Kaluza-Klein ansatz (\ref{kkans}). In particular, this implies that any elementary or solitonic $p$-brane solution is also a solution in the higher dimensions. However, it is important to realise that the resulting higher dimensional solution may not necessarily preserve the isotropic form of the $p$-brane ansatz (\ref{metrform}). In this paper, we are using the term `stainless' to describe the property of a brane solution of a lower-dimensional supergravity that cannot be oxidized into an isotropic brane solution in any supergravity in the next higher dimension.\footnote{We note that in defining a stainless $p$-brane to be one that cannot be oxidized to an isotropic brane in a higher dimension, we have not wanted to prejudge what a non-stainless $p$-brane may oxidize into. {\it A priori}, one could envisage that the extra dimension acquired upon oxidation could either become isotropically included into the world-brane dimensions, giving a $(p+1)$-brane in $(D+1)$ dimensions, {\it or} that the extra dimension could be isotropically included into the transverse dimensions, in which case one would still have a $p$-brane in the $(D+1)$ dimensions. The latter possibility, however, can never be realised within the scheme of Kaluza-Klein dimensional reduction because all fields are by construction taken to be independent of the extra coordinate, and this would be inconsistent with our ansatz (\ref{metrform}).} On the other hand, a $(p+1)$-brane solution in $(D+1)$ dimensions necessarily gives rise under dimensional reduction to an isotropic $p$-brane solution in $D$ dimensions. This automatic preservation of isotropicity for solutions under dimensional reduction corresponds directly to the process of double dimensional reduction \cite{dhis} of $p$-brane actions. The above ideas can be illustrated in our example of the bosonic Lagrangians (\ref{higherbos}) and (\ref{bosred}). First, we shall show that the elementary and solitonic solutions in $(D+1)$ dimensions reduce respectively to elementary and solitonic solutions in $D$ dimensions. In the case of an elementary solution, the $n$-index antisymmetric tensor in $(D+1)$ dimensions leads to an elementary brane with world volume dimension $\hat d =n-1$. The elementary ansatz for the $(D+1)$-dimensional field strength $\hat F_n$ in (\ref{higherbos}) is $\hat F_{m\mu_1\ldots\mu_{n-2} z} = \epsilon_{\mu_1\ldots\mu_{n-2} z} \partial_m e^C$. It follows from eqn (\ref{transgress}) that the corresponding $D$ dimensional fields $G_{n-1}$, ${G'}_n$ and ${\cal A}$ become \begin{eqnarray} G_{m\mu_1\ldots\mu_{n-2}} &=& \epsilon_{\mu_1\ldots\mu_{n-2}} \partial_m e^C \ ,\nonumber\\ {G'}_{\sst{M_1}\ldots\sst{M_n}} &=& 0\ ,\qquad {\cal A}_{\sst{M}}=0 \ .\label{elered} \end{eqnarray} This is nothing but the usual elementary-type ansatz for an $(n-1)$-index antisymmetric tensor in $D$ dimensions, and thus gives rise to an elementary brane solution (\ref{solution1}) with world volume dimension $d=n-2$. The metric ansatz in $(D+1)$ dimensions is given by $d\hat s^2 = e^{2\hat A}(dx^\mu dx^\nu \eta_{\mu\nu} + dz^2) + e^{2\hat B} dy^m dy^m$. In the elementary solution in $(D+1)$ dimensions, it follows from (\ref{solution1}) that $\phi= \hat a (D-1) \hat A/{\tilde d}$, and $\hat B= -(d+1) \hat A/{\tilde d}$. (Note that ${\tilde d}$ is the same for both $D$ and $(D+1)$ dimensions since, by definition, ${\tilde d} +2$ is the codimension of the world volume of the brane.) On the other hand in $D$ dimensions, we see from (\ref{bosred}) that the combination of scalar fields $-2(D-n)\alpha\varphi+ \hat a\phi=a\tilde\phi$, with $a^2=\hat a^2 +4(D-n)^2 \alpha^2$, defines the $SO(2)$-rotated $D$-dimensional dilaton $\tilde \phi$, whilst the orthogonal combination $2(D-n)\alpha \phi+ \hat a\varphi$ is set to zero. Since $n=d+2$, it then follows from (\ref{alphaval}) that $\hat a$ and $a$ are related by \begin{equation} \hat a^2 = a^2 - \fft{2{\tilde d}^2}{(D-2)(D-1)}\ .\label{arelation} \end{equation} Thus we find that $\tilde\phi=a (D-2) A/{\tilde d}$, $B=-d A/{\tilde d}$ and $e^{c A}= e^{\hat c\hat A}$, since, from the Kaluza-Klein ansatz (\ref{kkans}) for the metric, we have $\hat A=A+\alpha\varphi$ and $\hat B= B +\alpha\varphi$. But these expressions for $\tilde\phi$ and $B$ are precisely of the form given in (\ref{solution1}) for the elementary $(d-1)$-brane in $D$ dimensions. Thus we conclude that under dimensional reduction, an elementary $d$-brane in $(D+1)$ reduces to an elementary $(d-1)$-brane in $D$ dimensions. In the case of solitonic solutions, the analysis is parallel. The ansatz for the $n$-index antisymmetric tensor, which leads to a solitonic brane solution with world volume dimension $d=D-n$ in $(D+1)$ dimensions, takes the form $\hat F_{m_1\ldots m_n} = \lambda \epsilon_{m_1\ldots m_n p} \, y^p\, r^{-n-1}$. It follows from eqn (\ref{transgress}) that the corresponding $D$ dimensional fields ${G'}_n$, $G_{n-1}$ and ${\cal A}$ become \begin{eqnarray} {G'}_{m_1\ldots m_n} &=&\lambda \epsilon_{m_1\ldots m_n p}\, y^p\, r^{-n-1} \ ,\nonumber\\ G_{\sst{M_1}\ldots\sst{M_{n-1}}} &=& 0\ ,\qquad {\cal A}_{\sst{M}} = 0 \ . \end{eqnarray} This is indeed just the field configuration for a solitonic $(d-1)$-brane in $D$ dimensions. The analysis of the relation between the metrics in $(D+1)$ and $D$ dimensions is very similar to that in the elementary case. It is of interest to note that in the reduction of a $d$-brane in $(D+1)$ dimensions to a $(d-1)$ brane in $D$ dimensions, the degree of the antisymmetric tensor involved in the solution reduces by one in the elementary case, but remains unchanged in the solitonic case. Note also that the relation between $\hat a$ and $a$ in eqn (\ref{arelation}) is always satisfied in the dimensional reduction of a brane solution in $(D+1)$ to one in $D$ dimensions. This implies, conversely, that eqn (\ref{arelation}) is a {\it necessary} condition for the reverse procedure to be possible. It is easy to verify that the relation (\ref{arelation}) is uniquely satisfied with $\hat a$ and $a$ given by eqn (\ref{avalue}), provided that $\Delta$ is the same for both $\hat a$ and $a$. We have seen that brane solutions in higher dimensions can be reduced to those in lower dimensions {\it via} the Kaluza-Klein procedure; however, the inverse procedure is not necessarily possible. For example the $D$-dimensional bosonic Lagrangian (\ref{bosred}) that is derived from the $(D+1)$-dimensional Lagrangian (\ref{higherbos}) admits six brane solutions, namely an elementary and a solitonic solution for each of the three antisymmetric tensors $G_n$, $ G_{n-1}$ and $\cal F$. Two of these solutions are isotropically oxidizable to brane solutions in $(D+1)$ dimensions, by reversing the procedure discussed above, namely the elementary solution using $G_{n-1}$ and the solitonic solution using $G_n$. The remaining four solutions are stainless because they cannot be oxidized to isotropic brane solutions of the $(D+1)$ dimensional theory defined by eqn (\ref{higherbos}). To illustrate this, consider the elementary solution that uses the antisymmetric tensor $G_n$ in the $D$-dimensional Lagrangian (\ref{bosred}). The solution for the metric in $D$ dimensions is given by (\ref{metrform}) with $A$ and $B$ given in eqn (\ref{solution1}). This solution can be oxidized into a solution in $(D+1)$ dimensions, whose metric is given by \begin{equation} d\hat s^2 = e^{2\hat A} dx^\mu dx^\nu\eta_{\mu\nu} + e^{2\hat B} (dy^m dy^m + dz^2)\ ,\label{genoxiele} \end{equation} where \begin{equation} \hat A = \fft{(D-2)({\tilde d} +1)}{(D-1){\tilde d}} \, A\ ,\qquad \hat B = - \fft{(D-2)d}{(D-1){\tilde d}}\, A\ .\label{noname2} \end{equation} {}From the form of this $(D+1)$-dimensional metric, we can see that it does not describe an isotropic $d$-brane, since the different $r$-dependent prefactor for $dz^2$ prevents $z$ from being grouped together with the coordinates $x^\mu$. Note also that, although $dz^2$ does have the same prefactor as $dy^mdy^m$, this metric is still not isotropic in the transverse directions because the prefactors $e^{2\hat A}$ and $e^{2\hat B}$ are functions of $r=\sqrt{y^my^m}$ and not of $\sqrt{y^my^m + z^2}$. To summarise, we have seen that an elementary or solitonic $(p+1)$-brane solution in $(D+1)$ dimensions can always be reduced respectively to an elementary or solitonic $p$-brane solution in $D$ dimensions. On the other hand, the inverse process of dimensional oxidation to an isotropic brane solution is not always possible. Thus in a brane scan of elementary and solitonic solutions, we may factor out the rusty solutions and characterise the full solution set by the stainless $p$-branes only. There are three cases in which a $p$-brane solution can turn out to be stainless. The first case is when a brane solution arises in a supergravity theory that cannot be obtained by dimensional reduction, such as $D=11$ supergravity or type IIB supergravity in $D=10$. In the remaining two cases, the supergravity theory itself can be obtained by dimensional reduction, but oxidation to an isotropic brane solution is nonetheless not possible. In the second case, no $(D+1)$-dimensional supergravity theory has the necessary antisymmetric tensor for an isotropic brane solution. Specifically, if the $D$-dimensional solution is elementary, the $(D+1)$-dimensional theory would need an antisymmetric tensor of degree one higher than that in the $D$-dimensional theory. If it is instead a solitonic solution, the $(D+1)$-dimensional theory would need an antisymmetric tensor of the same degree as in the $D$-dimensional theory. In the third case, an antisymmetric tensor of the required degree exists in the $(D+1)$-dimensional theory, but the exponential dilaton prefactor has a coefficient $\hat a$ that does not satisfy eqn (\ref{arelation}). We shall meet examples of all three cases in the subsequent sections. \section{$D\ge10$ supergravity} $D=11$ is the highest dimension for any supergravity theory, and hence all the $D=11$ $p$-brane solutions are necessarily stainless. Since there is only one antisymmetric tensor field strength in the theory, namely a 4-index field, there is just one elementary membrane solution \cite{dust} and one solitonic 5-brane solution \cite{guv}. (Original papers giving $D=11$ supergravity, and all the other supergravities in various dimensions that we will consider here, can be found in \cite{SS}.) Dimensional reduction of $D=11$ supergravity to $D=10$ yields type IIA supergravity. The type IIA theory contains: a 2-form field strength giving rise to a particle and a 6-brane; a 3-form giving rise to a string and a 5-brane; and a 4-form giving rise to a membrane and a 4-brane. In each case we have listed first the elementary and then the solitonic solution. All of these solutions break half of the $D=10$, $N=2$ supersymmetry. Of the six solutions two, namely the elementary string and the solitonic 4-brane, can be oxidized to the corresponding elementary membrane and solitonic 5-brane in $D=11$. The remaining four solutions are stainless since $D=11$ supergravity lacks the necessary antisymmetric tensors. Note that the $11\longrightarrow 10$ situation corresponds precisely to the bosonic example we discussed in section 2.2. In addition, in $D=10$, there is the type IIB supergravity, which cannot be obtained by dimensional reduction from $D=11$. This theory contains a complex 3-form field strength giving rise to an elementary string and a solitonic 5-brane solution; and a self-dual 5-form field strength giving rise to a self-dual 3-brane \cite{dulu}. The string and 5-brane are in fact also solutions of $D=10$, $N=1$ supergravity, and are hence identical to the string and 5-brane solutions of the type IIA theory. Thus, although the type IIB theory cannot itself be obtained by dimensional reduction from $D=11$, these particular solutions of the IIB theory do have an oxidation pathway up to isotropic solutions in $D=11$. In such situations, we do not consider brane solutions to be stainless. The remaining solution, the self-dual 3-brane, is the only solution that belongs exclusively to the IIB theory. It is stainless and breaks half of the $N=2$ supersymmetry. \section{$D=9$ supergravity} \subsection{$N=1$, $D=9$ supergravity} $N=1$ supergravity in $D=9$ \cite{sgn} contains a 2-form field strength giving rise to an elementary particle and a solitonic 5-brane; and a 3-form field strength giving rise to an elementary string and a solitonic 4-brane. The solitonic 4-brane solution can be isotropically oxidized to the solitonic 5-brane of $N=1$, $D=10$ supergravity. The situation is somewhat more complicated for the oxidation of the elementary string solution. Obviously, this solution cannot be oxidized isotropically to an elementary membrane solution of $N=1$, $D=10$ supergravity because this theory lacks the necessary 4-form field strength, and thus no elementary membrane exists in the $N=1$, $D=10$ theory. Nonetheless, the $D=9$ string solution is not stainless because there is a different oxidation pathway available to it. The $D=9$ string can also be viewed as a solution of $N=2$, $D=9$ supergravity. In this guise, it can oxidize isotropically to a solution of type IIA $D=10$ supergravity, which {\it does} have a 4-form field strength. The elementary particle and solitonic 5-brane solution that arise from the 2-form field strength are stainless. Na\"\i vely, one might expect these solutions could oxidize up to the elementary string and solitonic 6-brane solutions of type IIA $D=10$ supergravity. However, as we showed in section 2.2, even when the necessary forms are present in the higher-dimensional theory an isotropic oxidation is possible only when the coefficient $a$ appearing in the dilaton prefactor $e^{-a\phi}$ satisfies the relation (\ref{arelation}). In the case of $N=1$, $D=9$ supergravity, the coefficient $a$ is given by eqn (\ref{avalue}) with $\Delta=2$. On the other hand, the coefficient $a$ in the type IIA, $D=10$ theory is given by eqn (\ref{avalue}) with $\Delta=4$. Since the $\Delta$ value has to be preserved under dimensional reduction, it follows that the particle and 5-brane solutions in $D=9$ are stainless. There {\it are} elementary particle and solitonic 5-brane descendants in $D=9$, nonetheless. These {\it are} obtained by dimensional reduction from the type IIA $D=10$ elementary membrane and solitonic 6-brane. From the $D=9$ point of view, these are obtained as solutions to $N=2$ supergravity using a 2-form field strength whose dilaton prefactor indeed has an $a$ coefficient given by (\ref{avalue}) with the necessary $\Delta=4$. The difference in $\Delta$ values establishes the distinctness of the stainless particle and 5-brane discussed above from those obtained by dimensional reduction. The metrics for the stainless elementary particle and solitonic 5-brane are given by \begin{eqnarray} {\rm elementary}:&& ds^2 = - \Big(1+\fft{k}{r^{6}}\Big)^{-12/7} dt^2 + \Big(1+\fft{k}{r^{6}}\Big)^{2/7} dy^mdy^m \ ,\nonumber\\ {\rm solitonic}:&& ds^2 =\Big (1+\fft{k}{r}\Big)^{-2/7} dx^\mu dx^\nu \eta_{\mu\nu} + \Big(1+\fft{k}{r}\Big)^{12/7} dy^mdy^m\ . \end{eqnarray} By contrast, the metrics for the elementary particle and solitonic 5-brane that can oxidize to an elementary string and a solitonic 6-brane in $D=10$ are given by \begin{eqnarray} {\rm elementary}:&& ds^2 = - \Big(1+\fft{k}{r^{6}}\Big)^{-6/7} dt^2 + \Big(1+\fft{k}{r^{6}}\Big)^{1/7} dy^mdy^m \ ,\nonumber\\ {\rm solitonic}:&& ds^2 = \Big(1+\fft{k}{r}\Big)^{-1/7} dx^\mu dx^\nu \eta_{\mu\nu} + \Big(1+\fft{k}{r}\Big)^{6/7} dy^mdy^m\ . \end{eqnarray} Let us now examine in detail the new stainless $D=9$ solutions. In particular, we need to verify that they preserve half of the supersymmetry. Since these solutions cannot be obtained from isotropic solutions in $D=10$, we do not have an automatic guarantee that half of the supersymmetry will be preserved. To investigate this, we first give the bosonic sector of the Lagrangian and the supersymmetry transformations. The bosonic sector of the Lagrangian is \begin{equation} {\cal L}=eR-\ft12e(\partial\phi)^2- \ft1{12}ee^{-\sqrt{\fft87}\phi}\, G_{\sst{MNP}}G^{\sst{MNP}}- \ft14ee^{-\sqrt{\fft27}\phi}\, F_{\sst{MN}}F^{\sst{MN}}\ , \label{d9lag} \end{equation} where $F_{\sst{MN}}=2\partial_{[\sst M}A_{\sst N]}$ and $G_{\sst{MNP}}=3\partial_{[\sst M}B_{\sst{NP}]}+\ft32 A_{[\sst M}F_{\sst{NP}]}$ \cite{sgn}. By comparison with eqn (\ref{avalue}) it is easy to verify that the $\Delta$ value for the 3-form $G$ is 4, but the value for the 2-form $F$ is 2. The supersymmetry transformation rules for the bosonic fields are: \begin{eqnarray} \delta e^{\sst A}{}_{\sst M} &=& -{\rm i} \, \bar\varepsilon\Gamma^{\sst A}\psi_{\sst M}\ ,\qquad \delta\phi = {\rm i}\sqrt2\, \bar\varepsilon\chi\ ,\nonumber\\ \delta A_{\sst M} &=& -\ft2{\sqrt{14}}\, e^{\sqrt{\fft1{14}}\phi}\, \bar\varepsilon\Gamma_{\sst M}\chi + \sqrt2 \, e^{\sqrt{\fft1{14}}\phi}\, \bar\varepsilon\psi_{\sst M}\ , \label{d9bostr}\\ \delta B_{\sst{MN}} &=& -2{\rm i}\, e^{\sqrt{\fft27}\phi}\, \bar\varepsilon\Gamma_{[\sst M}\psi_{\sst N]} + \ft{2{\rm i}}{\sqrt{7}}\, e^{\sqrt{\fft27}\phi}\, \bar\varepsilon \Gamma_{\sst{MN}}\chi + A_{[\sst M}\delta A_{\sst N]}\ .\nonumber \end{eqnarray} For the fermionic fields, the supersymmetry transformations are: \begin{eqnarray} \delta\chi &=& -\ft1{2\sqrt2}\Gamma^{\sst M}\varepsilon\, \partial_{\sst M}\phi +\ft1{12\sqrt7}\, e^{-\sqrt{\fft27}\phi}\, G_{\sst{MNP}} \Gamma^{\sst{MNP}}\varepsilon -\ft{{\rm i}}{4\sqrt{14}}\, e^{-\sqrt{\fft1{14}}\phi}\, F_{\sst{MN}} \Gamma^{\sst{MN}}\varepsilon\ ,\nonumber\\ \delta\psi_{\sst M} &=& D_{\sst M}\varepsilon + \ft1{84}e^{-\sqrt{\fft27}\phi}\, G_{\sst{NPQ}}\Big( \Gamma_{\sst M}{}^{\sst{NPQ}} - \ft{15}2\delta_{\sst M}^{\sst N}\, \Gamma^{\sst{PQ}}\Big)\varepsilon \nonumber\\ && -\ft{{\rm i}}{28\sqrt2}\, e^{-\sqrt{\fft1{14}}\phi}\, F_{\sst{NP}}\Big( \Gamma_{\sst M}{}^{\sst{NP}} - 12\delta_{\sst M}^{\sst N}\, \Gamma^{\sst P}\Big)\varepsilon\ .\label{d9fertr} \end{eqnarray} The elementary particle and solitonic 5-brane in $D=9$ dimensions are obtained from the ans\"atze for the 2-index antisymmetric tensor field strength $F_{\sst{MN}}$ given in (\ref{eleans}) and (\ref{solans}) respectively. The solutions are given by (\ref{solution1}). We shall first verify that the solitonic 5-brane solution preserves half of the supersymmetry. We begin by making a $6+3$ split of the gamma matrices: \begin{equation} \Gamma^{\mu} = \gamma^\mu \otimes \rlap 1\mkern4mu{\rm l}\ ,\qquad \Gamma^{m} = \gamma_7 \otimes \gamma^m \ ,\label{d95bs} \end{equation} where $\gamma_7 = \gamma_0\gamma_1\ldots\gamma_5$ on the world volume and $\gamma_1\gamma_2\gamma_3={\rm i}$ in the transverse space. Here, and throughout the paper, we adopt the convention that $\gamma_\mu$ and $\gamma_m$ are purely numerical matrices, with flat indices. The transformation rules for the fermionic fields in (\ref{d9fertr}) become \begin{eqnarray} \delta \chi &=& -\fft1{2\sqrt{2}}\, e^{-B}\, \partial_m \phi\, \gamma_7\otimes \gamma_m \varepsilon + \fft{\lambda}{2\sqrt{14}}\, e^{ -2B-\sqrt{\fft1{14}}\phi }\, \fft{y^m}{r^3}\, \rlap 1\mkern4mu{\rm l} \otimes \gamma_m \varepsilon \ ,\nonumber\\ \delta \psi_\mu&=& \fft12 \partial_m A\, e^{A-B}\, \gamma_\mu \gamma_7\otimes \gamma_m \varepsilon + \fft\lambda{14\sqrt2}\, e^{A-2B-\sqrt{\fft1{14}}\phi} \, \fft{y^m}{r^3}\, \gamma_\mu\otimes \gamma_m \varepsilon\ ,\label{d9soltr}\\ \delta \psi_m &=& \partial_m \varepsilon + \fft{{\rm i}}2 \partial_n B \varepsilon_{mnp} 1\otimes \gamma_p + \fft{\lambda}{14\sqrt2}\, e^{-B-\sqrt{\fft1{14}}\phi}\, \fft{y^m}{r^3}\, \gamma_7\otimes \rlap 1\mkern4mu{\rm l}\, \varepsilon \nonumber\\ &&- \fft{3{\rm i}\lambda}{7\sqrt{2}}\, e^{- B-\sqrt{\fft1{14}}\phi} \, \varepsilon_{mnp} \, \fft{y^n}{r^3}\, \gamma_7\otimes \gamma_p\, \varepsilon\ .\nonumber \end{eqnarray} Substituting the solitonic solution (\ref{solution1}) and noting that $\phi'$ and $A'$ satisfy (\ref{soluans}) and (\ref{betaeq}), we find that these variations all vanish provided that \begin{equation} \varepsilon = e^{\ft12 A} \, \varepsilon_0\ ,\qquad \gamma_7\otimes\rlap 1\mkern4mu{\rm l}\, \varepsilon_0 = \varepsilon_0\ ,\label{d9epso} \end{equation} where $\varepsilon_0$ is a constant spinor. Thus our solitonic 5-brane solution preserves half of the supersymmetry. We shall now verify that the elementary particle solution also preserves half the supersymmetry. We make a $1+8$ split of the gamma matrices: \begin{equation} \Gamma^0 = {\rm i} \gamma_9\ ,\qquad \Gamma^m = \gamma^m\ , \end{equation} where $\gamma_9=\gamma_1\gamma_2\cdots\gamma_8$. The transformation rules (\ref{d9fertr}) for the fermionic fields become \begin{eqnarray} \delta \chi &=& -\ft1{2\sqrt2}\, e^{-B} \partial_m \phi\, \gamma_m \varepsilon + \ft1{2\sqrt{14}}\, e^{-A - B +C- \sqrt{\fft1{14}}\phi}\, \partial_m C \,\gamma_m \gamma_9 \varepsilon\ ,\nonumber\\ \delta \psi_0 &=& \ft{\rm i}{2} e^{A-B}\, \partial_m A\, \gamma_m\gamma_9 \varepsilon -\ft{3{\rm i}}{7\sqrt2}\, e^{-B+C-\sqrt{\fft1{14}}\phi}\, \partial_m C \,\gamma_m \varepsilon\ ,\label{d9eletr}\\ \delta \psi_m &=& \partial_m \varepsilon +\ft{{\rm i}}2 \partial_n B\, \gamma_{mn} \varepsilon + \ft{1}{14\sqrt2}\, e^{-A +C-\sqrt{\fft1{14}}\phi}\, \partial_m C\, \gamma_{mn}\gamma_9 \varepsilon\nonumber\\ && -\ft3{7\sqrt{2}}\, e^{-A+C -\sqrt{\fft1{14}}\phi} \, \partial_m C \,\gamma_9 \varepsilon\ .\nonumber \end{eqnarray} Analogously to the solitonic case, the elementary particle solution also preserves half of the supersymmetry provided that \begin{equation} \varepsilon = e^{\ft12 A}\, \varepsilon_0\ ,\qquad \gamma_9\, \varepsilon_0 =\varepsilon_0\ ,\label{d9epel} \end{equation} So far we have obtained a stainless elementary particle and stainless solitonic 5-brane. Both solutions break half of the supersymmetry. The reason why these two solutions cannot be isotropically oxidized into $D=10$ dimensions is that both are obtained from the 2-index antisymmetric tensor field strength with the dilaton prefactor $e^{-a\phi}$ where the $a$ coefficient is given by (\ref{avalue}) with $\Delta = 2$, instead of the value $\Delta=4$ that characterises the prefactors of antisymmetric tensor field strengths in $D=10$. At first sight the occurrence of this new value of $\Delta$ may seem paradoxical since, Kaluza-Klein dimensional reduction preserves the value of $\Delta$, as we discussed for the scalar field $\tilde \phi$ defined below eqn (\ref{bosred}). Since $N=1$, $D=9$ supergravity can be obtained by dimensional reduction of $N=1$, $D=10$ supergravity, which has a single 3-form field strength, with a $\Delta = 4$ prefactor, it follows that all the antisymmetric tensors in $D=9$ will have $\Delta=4$ prefactors. The resolution of this apparent paradox involves details of the truncation of dimensionally reduced $N=1$, $D=10$ supergravity to the pure $N=1$ supergravity multiplet in $D=9$. The truncation removes a single $D=9$ Maxwell multiplet. The Lagrangian of the bosonic sector of $N=1$, $D=10$ supergravity is \begin{equation} {\cal L} = \hat e \hat R - \ft12 e (\partial\phi)^2 -\ft1{12} \hat e e^{-\phi} \, \hat F_{\sst{MNP}} \hat F^{\sst{MNP}}\ . \end{equation} Following the Kaluza-Klein dimensional reduction scheme discussed in section 2.2, this leads to the $D=9$ Lagrangian \begin{eqnarray} {\cal L} &=& e R -\ft12 e(\partial\phi)^2 -\ft12 e(\partial\varphi)^2 -\ft1{12} e e^{-\phi -\sqrt{\fft17}\varphi}\, {G'}_3^2 \nonumber\\ && -\ft14 e e^{-\fft4{\sqrt7}\varphi}\, {\cal F}^2 -\ft14 e e^{-\phi + \fft3{\sqrt7}\varphi}\, G_2^2\ .\label{n1d9lag} \end{eqnarray} As it stands, one cannot consistently truncate out either of the 2-form field strengths or either of the two scalars. Nonetheless, it is possible to make a consistent truncation to the bosonic sector of pure $N=1$, $D=9$ supergravity.\footnote{The possibility of making a consistent truncation to the $N=1$, $D=9$ supermultiplet may be shown using arguments similar to those in ref.\ \cite{zilch}} In order to do this, we must first rotate the basis for the scalar fields: \begin{equation} \phi = \sqrt{\ft78} \phi_1 -\sqrt{\ft18} \phi_2\ ,\qquad \varphi = \sqrt{\ft18} \phi_1 + \sqrt{\ft78} \phi_2\ . \end{equation} In terms of this rotated basis, the Lagrangian (\ref{n1d9lag}) becomes \begin{eqnarray} {\cal L} &=& e R -\ft12 e(\partial\phi_1)^2 -\ft12 e(\partial\phi_2)^2 -\ft1{12} e e^{-\sqrt{\fft87} \phi_1 }\, {G'}_3^2 \nonumber\\ && -\ft14 e e^{-\sqrt{\fft27} \phi_1 -\sqrt{2} \phi_2 }\, {\cal F}^2 -\ft14 e e^{- \sqrt{\fft27} \phi_1 +\sqrt{2}\phi_2}\, G_2^2\ .\label{n1d9lag2} \end{eqnarray} Now we can see that it is consistent with the equation of motion for $\phi_2$ to set $\phi_2=0$ provided that at the same time we set ${\cal F}$ equal to $G_2$. Defining then $F= \sqrt2{\cal F} = \sqrt2 G_2$, we obtain the Lagrangian for the bosonic sector of pure $N=1$, $D=9$ supergravity: \begin{equation} {\cal L} = e R -\ft12 e(\partial \phi_1)^2 -\ft1{12} e e^{-\sqrt{\fft87}\phi_1} {G'}_3^2 -\ft14 e e^{-\sqrt{\fft27}\phi_1} F^2\ ,\label{n1d9sez} \end{equation} where $d {G'}_3 + \ft12 F\wedge F =0$. This result coincides with the Lagrangian given in ref.\ \cite{sgn}, which appears in eqn (\ref{d9lag}). Thus although the value of $\Delta$ for the combinations $- \sqrt{\ft27} \phi_1 \pm \sqrt{2}\phi_2$ occurring in the 2-form field strength prefactors before truncation is $\Delta=4$, the value after the truncation in which $\phi_2$ is set equal to zero is $\Delta = 2 $.\footnote{ Another point of view for resolving the apparent paradox is to regard the stainless particle and 5-brane as solutions of the full dimensionally reduced $N=1$, $D=10$ supergravity, {\it i.e.}\ $N=1$, $D=9$ supergravity plus the Maxwell multiplet. From this point of view, these solutions fall outside out $p$-brane ans\"atze (\ref{eleans}) and (\ref{solans}) because more than one antisymmetric tensor field strength takes a non-vanishing value. The solutions arising from this new ansatz are equivalent to those in the truncated $N=1$ theory with $\Delta = 2$.} Having studied this example in detail, we are now in a position to be more precise about the possible values of $\Delta$ that can arise in supergravity theories. We have seen that we may treat $D=11$ supergravity, which has no dilaton, as having the value $\Delta=4$ for its 4-form field strength, since this value corresponds, by virtue of eqn (\ref{avalue}), to $a=0$. We have also seen that pure Kaluza-Klein dimensional reduction, where one performs no truncation on the lower-dimensional theory, preserves the values of $\Delta$ from the higher dimension. Thus in the absence of any truncation, all supergravity theories that are obtained by dimensional reduction from $D=11$ will have $\Delta=4$ for all dilaton couplings. However, as we demonstrated in the case of $N=1$, $D=9$ supergravity above, if a supergravity theory in a lower dimension is obtained by a process of {\it truncation} as well as dimensional reduction, then the values of $\Delta$ for the coupling of the particular combinations of dilaton fields that survive the truncation to the antisymmetric tensor combinations that survive the truncation can differ from 4. For example, one can have $\Delta=2$ for 2-form field strengths in $D\le9$ supergravities. Before ending this section, it is of interest to investigate the warped metrics that one does obtain in $D=10$ if one oxidizes the stainless elementary particle and solitonic 5-brane from $D=9$, so as to compare them with the isotropic metrics of the elementary string and and solitonic 6-brane occurring in $D=10$. The metrics obtained by oxidizing the stainless $D=9$ solutions are given by \begin{eqnarray} {\rm elementary:}&& d\hat s^2 = \Big ( 1 + \fft{k}{r^6}\Big )^{-7/4} dx^\mu dx^\nu\eta_{\mu\nu} + \Big (1+ \fft{k}{r^6} \Big)^{1/4} \Big (dy^m dy^m + (dz + {\cal A})^2\Big)\ , \nonumber\\ {\rm solitonic:} && d\hat s^2 = \Big ( 1 + \fft{k}{r}\Big )^{-1/4} \Big( dx^\mu dx^\nu\eta_{\mu\nu} + (dz + {\cal A})^2\Big) + \Big (1+ \fft{k}{r} \Big)^{7/4} dy^m dy^m\ .\label{d9oxid} \end{eqnarray} Here we see that we have pushed oxidation too far: neither of these two metrics describes isotropic brane solutions in $D=10$. In both cases there is a non-vanishing gauge potential ${\cal A}={\cal A}_{\sst M} dx^{\sst M}$, which describes a topologically non-trivial field configuration, implying that $z$ is a coordinate on a non-trivial $U(1)$ fibre bundle, and thus the metric is `twisted.' Furthermore, in order for this coordinate to be well defined, it must be taken to be periodic with period $\Delta z = \int {\cal F}$ (or $\int{\cal F}$ divided by any integer). In the elementary case, as we also saw in the general example given in eqn (\ref{genoxiele}), the metric would not be isotropic even if ${\cal A}$ were equal to zero, for the reasons we discussed. By contrast, the metrics for the {\it isotropic} elementary string and solitonic 6-brane are given by (\ref{metrsol}) with $D=10$ and $\Delta = 4$, by taking $d=2$ and $d=7$ respectively. \subsection{$N=2$, $D=9$ supergravity} $N=2$ supergravity in $D=9$ contains three 2-form, two 3-form and one 4-form field strengths. In addition there are three scalar fields. Two of these behave like dilatons and appear undifferentiated in exponential prefactors multiplying the kinetic terms for the antisymmetric tensors. The third scalar does not appear in exponential prefactors in the Lagrangian; furthermore, its kinetic term itself has a dilaton prefactor. Thus we may view this scalar field as the 0-form potential for a 1-form field strength. We can use this field strength to obtain a solitonic 6-brane in $D=9$. $N=2$, $D=9$ supergravity has not yet been constructed; however, it could be easily obtained by dimensional reduction of type IIA supergravity in $D=10$. We expect that the elementary and solitonic brane solutions that are obtained from the 2-form, 3-form and 4-form field strengths are either obtainable by dimensional reduction from those in $D=10$ or are equivalent to the stainless solutions we constructed in $N=1$, $D=9$ supergravity. However, the solitonic 6-brane that is associated with the 1-form field strength is necessarily stainless, since the 1-form field strength appears first in $D=9$ supergravity in the descent from eleven dimensions. We shall first obtain the solution and then shall verify that it preserves half of the supersymmetry. The Lagrangian of the relevant part of the bosonic sector of $N=2$, $D=9$ supergravity can be obtained by Kaluza-Klein dimensional reduction of the metric, dilaton and 2-form field strength in type IIA supergravity in $D=10$, whose Lagrangian is given by eqn (\ref{higherbos}) with $n=2$ and $\hat a = 3/2$. The reduced $9$-dimensional Lagrangian is given by (\ref{bosred}), again with $n=2$ and $\hat a=3/2$. In order to obtain a solitonic 6-brane solution, we can consistently set ${\cal F}=0$, ${G'}_2=0$ and furthermore truncate out one of the two scalar field degrees of freedom by setting \begin{equation} \ft32 \phi - 14\alpha \varphi = 2 \tilde \phi\ ,\qquad 14\alpha \phi + \ft32 \varphi = 0\ . \end{equation} Thus the Lagrangian for the relevant bosonic fields in $D=9$ is \begin{equation} {\cal L} = e R - \ft12 e (\partial \tilde\phi)^2 - \ft12 e e^{-2\tilde \phi} \, G_1^2\ . \end{equation} This construction, which precisely parallels the previous discussions for general values of $n$, emphasises that $G_{\sst M}=\partial_{\sst{M}}b$ should properly be thought of as the field strength for the 0-form gauge potential $b=\hat A_z$, since it has its origin in the gauge field $\hat F_2$ in $D=10$. Thus it is legitimate for $G_1$ to take the necessary topologically non-trivial form in the solitonic 6-brane solution in $D=9$, in which its 0-form potential is well-defined only in patches. Using the ansatz for $G_1$ given by eqn (\ref{solans}), we can obtain the 6-brane solution. However in this case ${\tilde d}=0$, and hence the general solution given by eqn (\ref{solution1}) no longer applies. Nonetheless the equations (\ref{eqmo2}) are easy to solve; the metric of the solitonic 6-brane in $D=9$ is given by \begin{equation} ds^2 = dx^\mu dx^\nu \eta_{\mu\nu} + (1 + k \log{r}) \, dy^m dy^m\ , \end{equation} and the dilaton field $\tilde \phi$ is given by $e^{\tilde \phi} = 1+ k \log r$. It satisfies $\tilde\phi'=S$, where $S$ is given in eqn (\ref{sforms}). If $N=2$, $D=9$ supergravity had been constructed, it would have been a simple matter to check whether the above solution preserved half of the supersymmetry. In lieu of this, we may exploit the fact that Kaluza-Klein dimensional reduction preserves unbroken supersymmetry, and carry out the computation for the corresponding oxidized brane solution in $D=10$. Of course, since the 6-brane in $D=9$ is stainless, the resulting oxidized metric will not be an isotropic 7-brane. In fact it takes the form \begin{equation} d\hat s^2 = e^{-\fft18 \tilde\phi}\, dx^\mu dx^\nu \eta_{\mu\nu} + e^{\fft78 \tilde\phi}\, \Big (dy^m dy^m + dz^2\Big)\ .\label{ox6brane} \end{equation} The relevant terms in the fermionic transformation rules of type IIA, $D=10$ supergravity, involving the non-vanishing 2-form field strength $\hat F_2$, are: \begin{eqnarray} \delta \chi &=& \ft{\sqrt2}{4} \partial_{\hat\sst{M}}\phi\, \hat\Gamma^{\hat \sst{M}} \hat\Gamma^{11}\varepsilon -\ft{3}{16\sqrt 2} e^{-\fft34\phi}\, \hat F_{\hat\sst{M}\hat\sst{N}} \hat\Gamma^{\hat\sst{M}\hat\sst{N}} \varepsilon\ ,\nonumber\\ \delta \psi_{\hat\sst{M}} &=& \hat D_{\hat\sst{M}}\varepsilon -\ft1{64} e^{-\fft34 \phi}\, \hat F_{\hat\sst{N}\hat\sst{P}}\Big( \hat\Gamma_{\hat\sst{M}}{}^{ \hat\sst{N}\hat\sst{P}} -14 \delta_{\hat\sst{M}}^{\hat\sst{N}}\, \hat\Gamma^{\hat\sst{P}}\Big)\hat\Gamma^{11}\varepsilon \ .\label{d10susy} \end{eqnarray} It follows from (\ref{transgress}) that the 2-form field strength is given by $\hat F_{mz} = G_m = \lambda\varepsilon_{mn}\, y^n/r^2$. The functions $A$ and $B$ appearing in the $D=9$ solitonic 6-brane metric, the Kaluza-Klein scalar $\varphi$, and the $D=10$ dilaton $\phi$ are given in terms of $\tilde\phi$ by $A=0$, $B=\ft12\tilde\phi$, $\alpha\varphi=-\ft1{16}\tilde\phi$, $\phi=\ft34\tilde\phi$. The Kaluza-Klein vector potential ${\cal A}_{\sst M}$ is equal to zero. With these, and the expressions (\ref{spincon}) for the $D=10$ spin connection appearing in $\hat D_{\hat\sst{M}}$ in terms of the $D=9$ spin connection and $\varphi$, it is now straightforward to substitute the oxidized solution into the fermionic transformation rules given in eqn (\ref{d10susy}). We find that half the supersymmetry is preserved if $\varepsilon$ satisfies the conditions \begin{equation} \varepsilon = e^{-\ft1{32}\tilde\phi}\, \varepsilon_0\ ,\qquad \hat\Gamma_{{\underline m}{\underline n}}\, \varepsilon_0 = - \epsilon_{mn}\, \hat\Gamma_{\underline z}\, \hat\Gamma^{11}\, \varepsilon_0\ , \end{equation} where $\varepsilon_0$ is a constant spinor. Having demonstrated in $D=10$ that the non-isotropic oxidation (\ref{ox6brane}) of the 6-brane preserves half of the type IIA, $D=10$ supersymmetry, it follows that the 6-brane solution itself in $D=9$ also preserves half of the $N=2$, $D=9$ supersymmetry. \section{$D\le 8$ supergravity} As one descends through the dimensions, starting at $D=11$, one encounters various stainless brane solutions. First of all, they occur if the supergravity theory in a given dimension cannot be obtained from dimensional reduction. This happens in $D=11$, and $D=10$ for type IIB supergravity. A second reason for the occurrence of stainless brane solutions is if no supergravity theory in the next higher dimension has the necessary antisymmetric tensor field strength. The above two reasons account for all the stainless brane solutions in $D=11$ and $D=10$, and the stainless solitonic 6-brane in $D=9$. By the time one has reached $D=9$, all possible degrees $n\le D/2$ for antisymmetric tensor field strengths have occurred. Because of this, any further stainless brane solutions in $D\le 8$ will arise only for the third of the reasons we discussed in section 2.2, namely, that the $\Delta$ values for the exponential dilaton prefactors of the relevant antisymmetric tensors in the higher and lower dimensions differ. This phenomenon already occurred for the 2-form field strength in $D=9$, giving rise to the stainless elementary particle and solitonic 5-brane, as we discussed in the previous section. In view of the above considerations, it is not surprising that further stainless brane solutions in $D\le 8$ are relatively sparse. However, we shall not attempt in this paper to give a full classification of the super $p$-brane solutions for $D\le 8$. In $D=8$ and $D=7$, there are stainless elementary particle solutions. These solutions arise using the 2-form field strength with $\Delta = 2$. They are stainless since all the 3-form field strengths in one dimension higher have $\Delta = 4$. In $D=6$, analogously to the cases of $D=8$ and $D=7$, there is also a stainless elementary particle obtained from the 2-form field strength with $\Delta =2$, which is part of the supergravity multiplet in $N=2$, $D=6$ supergravity. In $N=1$, $D=6$ supergravity, on the other hand, there exists a self-dual 3-form field strength, and there is no dilaton. As we discussed in section 2.2, brane solutions are still given by eqn (\ref{solution1}), with $a$ set to zero, even in the absence of the dilaton. Thus this self-dual string solution is equivalent to the case where $\Delta = 2$, with the metric given by (\ref{metrsol}). Since there is no supergravity theory in $D=7$ that contains a 3-form or 4-form field strength with $\Delta=2$, the self-dual string in $D=6$ is stainless. The existence of a 3-form with $\Delta=2$ in $D=6$ implies that there is no further stainless elementary particle in $D\le 5$ that arises from the 2-form field strength with $\Delta = 2$. However, in $N=1$, $D=5$ supergravity, a new value of $\Delta$ for the 2-form field strength arises, namely $\Delta = 4/3$. This reflects the fact that there is no dilaton in the theory. This 2-form field strength accordingly gives rise to a stainless elementary particle and a stainless solitonic string \cite{GHT}, with metrics given by eqn (\ref{metrsol}). To see how this works, we can carry out the Kaluza-Klein dimensional reduction of $N=1$, $D=6$ supergravity. Its bosonic sector comprises just the metric tensor and the self-dual 3-form field strength mentioned above. Since there is no covariant action for this theory, we must instead implement the dimensional reduction on the equations of motion themselves. The bosonic equations of motion are given by \begin{equation} \hat R_{\hat\sst{M} \hat\sst{N}} = \ft14 \hat F_{\hat\sst{M}\hat\sst{P} \hat\sst{Q}}\hat F_{\hat\sst{N}}{}^{\hat\sst{P}\hat\sst{Q}}\ ,\qquad {\hat F}_{\hat\sst{M}\hat\sst{N}\hat\sst{P}}= *{\hat F}_{\hat\sst{M}\hat\sst{N}\hat\sst{P}} \ .\label{d6eqmo} \end{equation} The Kaluza-Klein ansatz for the metric and the antisymmetric tensor are given by (\ref{kkans}) and (\ref{transgress}) as usual, but now, the self-duality condition $\hat F= *\hat F$ implies that the lower-dimensional 2-form and 3-form field strengths $G_2$ and ${G'}_3$ are related: \begin{equation} {G'}_{\sst{ABC}}= \ft12 e^{4\alpha\varphi}\, \epsilon_{\sst{ABCDE}}\, G^{\sst{DE}}\ , \end{equation} where $\alpha$, given by (\ref{alphaval}), takes the value $\alpha=1/(2\sqrt6)$. Substituting these ans\"atze into the 6-dimensional equations of motion (\ref{d6eqmo}), and making use of the expressions (\ref{ricten}) for the Ricci-tensor components, we obtain the 5-dimensional equations of motion \begin{eqnarray} R_{\sst{AB}} &=& \ft12 \partial_{\sst{A}}\varphi\, \partial_{\sst{B}}\varphi + \ft12 e^{-8\alpha\varphi}\, ({\cal F}^2_{\sst{AB}} -\ft16 {\cal F}^2 \eta_{\sst{AB}}) + e^{4\alpha\varphi}\, (G^2_{\sst{AB}} -\ft16 G^2 \eta_{\sst{AB}})\ ,\nonumber\\ \nabla^{\sst{B}}\Big( e^{-8\alpha\varphi}\, {\cal F}_{\sst{AB}} \Big) &=& \ft14 e^{3\alpha\varphi} \epsilon_{\sst{ABCDE}} G^{\sst{BC}} G^{\sst{DE}}\ ,\label{d5eqmo}\\ \mathord{\dalemb{6.8}{7}\hbox{\hskip1pt}} \varphi &=& 2\alpha e^{4\alpha\varphi}\, G^2 - 2\alpha e^{-8\alpha \varphi}\, {\cal F}^2\ .\nonumber \end{eqnarray} We see that we may consistently truncate these fields to those of minimal $D=5$ supergravity, whose bosonic sector comprises just the metric and a 2-form field stength, by setting $\varphi=0$ and ${\cal F}_{\sst{AB}} =G_{\sst{AB}}$. Defining $F_{\sst{AB}}\equiv \sqrt3{\cal F}_{\sst{AB}}= \sqrt3 G_{\sst{AB}}$, we find that the equations of motion for the remaining fields can be derived from the Lagrangian \begin{equation} {\cal L} =e R -\ft14 e F^2 -\ft1{12\sqrt3}\epsilon^{\sst{MNPQR}}\, F_{\sst{MN}} F_{\sst{PQ}} A_{{\sst R}}\ ,\label{d5lag} \end{equation} where $F_{\sst{MN}}=2\partial_{[\sst{M}} A_{\sst{N}]}$. This Lagrangian describes the bosonic sector of minimal $D=5$ supergravity. We see that a 2-form field strength with a new value of $\Delta$ has emerged in the descent to five dimensions, namely $\Delta=\ft43$ and hence $a=0$. It follows that brane solutions in minimal $D=5$ supergravity, which make use of this 2-form field strength, cannot be oxidized to give isotropic brane solutions in any higher dimension. In this way, we obtain the stainless elementary particle and solitonic string solutions referred to above. Their metrics are given by (\ref{metrsol}), with $d=1$, ${\tilde d}=2$ and $d=2$, ${\tilde d}=1$ respectively, where $\Delta=\ft43$. To check the unbroken supersymmetry of these solutions, we need the gravitino transformation rule in $D=5$ simple supergravity, which reads \begin{equation} \delta \psi_{\sst M} = D_{\sst M}\varepsilon -\ft{{\rm i}}{8\sqrt3} F_{\sst{NP}}\Big(\Gamma_{\sst M}{}^{\sst{NP}} -4 \delta_{\sst M}^{\sst N} \Gamma^{\sst P} \Big)\varepsilon \ .\label{d5susy} \end{equation} For the solitonic string, we decompose the $D=5$ gamma matrices in the $2+3$ split $\Gamma^\mu=\gamma^\mu\otimes \rlap 1\mkern4mu{\rm l}$, $\Gamma^m=\gamma_3 \otimes \gamma^m$, where $\gamma_0 \gamma_1 =\gamma_3$ on the brane volume, and $\gamma_1\gamma_2\gamma_3={\rm i}$ in the transverse space. Thus we have \begin{eqnarray} \delta\psi_\mu &=& \fft12 \partial_m A\, e^{A-B}\, \gamma_\mu\gamma_3\otimes \gamma_m \varepsilon +\fft{\lambda}{4\sqrt3}\, e^{A-2B}\, \fft{y^m}{r^3}\, \gamma_\mu\otimes \gamma_m \varepsilon\ ,\nonumber\\ \delta\psi_m &=& \partial_m\varepsilon +\fft{{\rm i}}{2}\partial_n B\, \epsilon_{mnp} \rlap 1\mkern4mu{\rm l}\otimes\gamma_p \varepsilon + \fft{\lambda}{4\sqrt3} e^{-B}\, \fft{y^m}{r^3}\, \gamma_3\otimes\rlap 1\mkern4mu{\rm l}\, \varepsilon -\fft{{\rm i} \lambda}{2\sqrt3}e^{-B}\, \fft{y^n}{r^3}\, \gamma_3\otimes\gamma_p\, \varepsilon \ .\label{d5solsusy} \end{eqnarray} Substituting the solitonic string solution, which, from (\ref{abphirel}) in the limit $a=0$ has $A'=-\ft12 B'= -\lambda/(2\sqrt3) e^{-B}r^{-2}$, we find that half of the supersymmetry is preserved provided that \begin{equation} \varepsilon =e^{\ft12 A}\, \varepsilon_0\ , \qquad \gamma_3\otimes\rlap 1\mkern4mu{\rm l} \, \varepsilon_0=\varepsilon_0\ ,\label{stringeps} \end{equation} where $\varepsilon_0$ is a constant spinor. For the elementary particle, we decompose the $D=5$ gamma matrices in the $1+4$ split $\Gamma^0={\rm i} \gamma_5$, $\Gamma^m= \gamma^m$, where $\gamma_5=\gamma_1\gamma_2\gamma_3\gamma_4$. The supersymmetry transformation rule (\ref{d5susy}) becomes \begin{eqnarray} \delta\psi_0 &=& \ft{{\rm i}}{2} e^{A-B}\,\partial_m A\, \gamma_m\gamma_5\varepsilon -\ft{{\rm i}}{2\sqrt3} e^{C-B}\, \partial_m C\, \gamma_m \varepsilon\ ,\nonumber\\ \delta\psi_m &=& \partial_m\varepsilon +\ft12 \partial_n B\, \gamma_{mn}\varepsilon +\ft{1}{4\sqrt3} e^{C-A}\, \partial_n C\, \gamma_{mn}\gamma_5\varepsilon -\ft{1}{2\sqrt3} e^{C-A}\, \partial_m C\, \gamma_5\varepsilon\ . \label{d5susyele} \end{eqnarray} Substituting the elementary particle solution, which, from (\ref{abphirel}) in the limit $a=0$ has $A'=-2B'= (1/\sqrt3) e^{C-A}\, C'$, we find that half of the supersymmetry is preserved provided that \begin{equation} \varepsilon=e^{\ft12 A}\, \varepsilon\ ,\qquad \gamma_5\, \varepsilon_0= \varepsilon_0\ , \end{equation} where $\varepsilon_0$ is a constant spinor. It is interesting to note that there are in total {\it three} inequivalent solitonic string solutions in $D=5$, namely the stainless example we have just derived, a rusty string that oxidizes to our stainless 5-brane in $D=9$, and another rusty string that oxidizes to the stainless 6-brane in $D=10$. Their metrics in $D=5$ are given by (\ref{metrsol}) with $d=2$ and ${\tilde d}=1$, by taking $\Delta=\ft43$, $\Delta=2$ and $\Delta=4$ respectively. Upon dimensional reduction to $D=4$, they give rise to particles with $a=1/\sqrt3$, 1 and $\sqrt 3$ respectively. These correspond to the black hole solutions of $D=4$ string theory (see, for example, \cite{dkl}).\footnote{We are grateful to J. Rahmfeld for drawing our attention to the black hole solutions in the $D=4$ string.} \section{Zero modes} In the previous sections, we described stainless $p$-brane solutions in various dimensions. The complete set of brane solutions is thus given by those solutions together with their descendants {\it via} Kaluza-Klein double dimensional reduction. All of these solutions break half of the supersymmetry. Each broken supersymmetry transformation in a $p$-brane solution gives rise to a corresponding fermionic Goldstone zero mode. There will also be bosonic zero modes associated with the breaking of local bosonic gauge symmetries by the non-vanishing $p$-brane background solution. These will certainly include the translational zero modes corresponding to the broken constant general coordinate transformations $\delta y^m=c^m$ in the space transverse to the $p$-brane world volume. Thus there will be $D-d={\tilde d}+2$ such bosonic zero modes. Since supersymmetry remains partially unbroken by the solution, it follows that the fermionic and bosonic zero modes must form supermultiplets under the remaining unbroken supersymmetry. In particular, there must be equal numbers of fermionic and bosonic zero-mode degrees of freedom. The matching of the zero modes for the bosonic and fermionic fields is straightforward in the case of the elementary membrane in $D=11$ and the solitonic 5-brane in $D=10$, and also for all their descendants {\it via} dimensional reduction. In all of these cases, the number of translational zero modes is precisely the same as that of the fermionic zero modes, {\it i.e.}\ the number of on-shell fermionic zero-mode degrees of freedom. Thus for the supermembrane in $D=11$, there are $8=32/2/2$ fermionic zero modes, where the original 32 components of the supersymmetry parameter in $D=11$ are halved once to arrive at the number of on-shell degrees of freedom, and halved again because half of the supersymmetries are broken. The membrane solution breaks translational invariance in the $y^m$ directions, giving rise to $8=11-3$ bosonic zero modes. The same counting of $8+8$ degrees of freedom holds for the dimensional reduction to the string in $D=10$. For the solitonic 5-brane in $N=1$, $D=10$ supergravity, there are $4=16/ 2/2$ fermionic zero modes, and $4=10-6$ bosonic translational zero modes. This matching of $4+4$ degrees of freedom holds for the various stages of dimensional reduction all the way down to the string in $D=6$ and the superparticle in $D=5$. In all the other brane solutions, the number of translational zero modes is less than the number of fermionic zero modes. Since we know that the remaining unbroken symmetry guarantees a matching of the bose and fermi zero modes, it follows that the there must be further bosonic zero modes associated with these solutions. They arise from the breaking of antisymmetric tensor gauge symmetries. The simplest way to find these additional bosonic zero modes is first to construct the fermionic zero modes, and then to obtain their bosonic partners by transforming them under the remaining unbroken supersymmetries. We shall carry out this procedure first for the stainless solitonic 5-brane in $N=1$, $D=9$ supergravity, which we constructed in section 4.1. This solution has four fermionic zero modes; however, it has only $9-6=3$ translational zero modes. As we shall see, there is one further bosonic zero mode associated with the breaking of the gauge invariance of the 2-form field strength that takes a non-zero value in the background solution. The fermionic supersymmetry transformations in the background of the solitonic 5-brane are given by eqn (\ref{d9soltr}). As we discussed in section 4.1, these variations vanish for spinors $\varepsilon$ satisfying (\ref{d9epso}), which includes a chirality condition. They correspond to the unbroken supersymmetry generators. The broken generators, on the other hand, correspond to supersymmetry parameters $\eta$ that have the opposite chirality under the $\gamma_7$ matrix on the world volume. Specifically, we shall consider spinors $\eta$ given by \begin{equation} \eta= e^{-\ft12 A}\, \eta_0\ ,\qquad \gamma_7\otimes\rlap 1\mkern4mu{\rm l}\, \eta_0 =-\eta_0\ , \label{etasol} \end{equation} where $\eta_0$ is constant. This choice is motivated by the simplifications to the fermionic zero-mode structure that result. Note that any other asymptotically constant spinors with the same $\gamma_7$ eigenvalue could equally well have been chosen. These would lead to zero-modes differing from ours by pure gauge transformations whose parameters die off at infinity. With our choice, it follows from (\ref{d9soltr}) that the purely bosonic 5-brane soliton background varies into the following fermionic configuration: \begin{eqnarray} \chi&=& \ft{1}{\sqrt2} e^{-\ft12 A-B} \, \partial_m\phi\, \rlap 1\mkern4mu{\rm l}\otimes\gamma_m\, \eta_0\ ,\nonumber\\ \psi_\mu&=& -\partial_m A\, e^{\ft12 A -B}\,\gamma_\mu\otimes\gamma_m\, \eta_0 \ , \label{d95brfermzm}\\ \psi_m&=& \partial_n B\, e^{-\ft12 A}\, \rlap 1\mkern4mu{\rm l}\otimes\gamma_{mn} \eta_0\ .\nonumber \end{eqnarray} These, then, describe the four fermionic zero modes, parametrised by the eight independent spinors $\eta_0$ that satisfy the chirality condition given in (\ref{etasol}). (Recall that for $d\ge3$, the count of on-shell fermionic degrees of freedom is half that of the off-shell spinor fields.) We can substitute these spinors into the bosonic transformation rules (\ref{d9bostr}), taking the supersymmetry parameter $\varepsilon$ to be one of the eight unbroken generators given by (\ref{d9epso}), in order to obtain the bosonic superpartners of the fermionic zero modes. Carrying out this procedure, we find the following non-vanishing results for the bosonic zero modes: \begin{eqnarray} \delta\phi &=& \delta_{\rm Diff}\, \phi\, \nonumber\\ \delta e^{\underline\mu}{}_\nu &=& \delta_{\rm Diff}\, e^{\underline\mu}{}_\nu + \Omega^{\underline\mu}{}_{\underline\rho} \, e^{\underline \rho}{}_\nu\ ,\nonumber\\ \delta e^{\underline m}{}_n &=& \delta_{\rm Diff}\, e^{\underline m}{}_n + \Omega^{\underline m}{}_{\underline p} \, e^{\underline p}{}_n\ ,\label{zmodes}\\ \delta A_m &=& \delta_{\rm Diff}\, A_m + \partial_m\widetilde\Lambda\ , \nonumber\\ \delta B_{mn} &=& \delta_{\rm Diff}\, B_{mn} -\ft12 \widetilde\Lambda F_{mn} + \partial_{[m} \widetilde\Lambda_{n]}\ ,\nonumber \end{eqnarray} where $\delta_{\rm Diff}$ denotes a diffeomorphism transformation, $\delta_{\rm Diff}\, V_m = c^n\partial_n V_m + \partial_m c^n\, V_n$, {\it etc}., and the composite transformation parameters are given by\footnote{These results could be obtained from the commutator algebra of the local supersymmetry transformations. To see this, note that the bosonic zero modes $b$ can be written in terms of the femionic zero modes $f$ as $b=\delta_\epsilon f$. However, $f$ can be written as $f=\delta_\eta B$, where $B$ represents the supersymmetric bosonic background fields. Since $\delta_\epsilon B=0$, we can write $b=[\delta_\epsilon,\delta_\eta] B= \delta_c B+\delta_{\Lambda_{0}} B+\delta_{\Lambda_1} B+\delta_\Omega$, where $c$, $\Lambda_{0}$, $\Lambda_{1}$ and $\Omega$ are the composite parameters for the general coordinate, abelian gauge, antisymmetric gauge and Lorentz transformations, respectively.} \begin{eqnarray} c^m &\equiv& {\rm i} e^{-B} {\bar \varepsilon}_0 \rlap 1\mkern4mu{\rm l}\otimes \gamma_m\eta_0\ , \nonumber\\ \Lambda &\equiv& \sqrt2 e^{-A}\, {\bar \varepsilon}_0\eta_0\ ,\nonumber\\ \Lambda_m &\equiv& -\ft74 e^{-2A}\, c_m\ ,\label{zmodeparms}\\ \widetilde\Lambda &\equiv& \Lambda -c^m A_m\ , \qquad \widetilde \Lambda_m \equiv \Lambda_m - \Lambda A_m + 2 c^p B_{mp}\ \nonumber\\ \Omega^{\underline \mu}{}_{\underline\rho} &\equiv& {\rm i} \bar\varepsilon_0 \gamma^{\mu}{}_\rho\otimes \gamma_m \eta_0 \,e^{-B}\, \partial_m A\ ,\nonumber\\ \Omega^{\underline m}{}_{\underline p} &\equiv& \partial_m c^p - \partial_p c^m -\epsilon^m{}_{pq}\, \bar\varepsilon_0 \eta_0 \,e^{-B}\,\partial_q B\ . \nonumber \end{eqnarray} The quantities $\widetilde\Lambda$ and $\widetilde\Lambda_m$ are gauge transformation parameters for the potentials $A_m$ and $B_{mn}$ respectively; $\widetilde\Lambda$ appears also in the $B_{mn}$ transformation because its field strength is given by $G=dB + \ft12 A\wedge F$. As one chooses different constant spinors $\varepsilon_0$ and $\eta_0$, satsifying their respective $\pm1$ chirality conditions under $\gamma_7$, the diffeomorphism parameter $c^m$ and gauge transformation parameter $\Lambda_m$ jointly fill out a 3-dimensional space of gauge transformations that are constant and non-vanishing as $r$ tends to infinity. Likewise, $\Lambda$ describes a further independent asymptotically constant non-vanishing gauge transformation. Taken together, we have the four independent bosonic zero modes of the solitonic 5-brane in $D=9$ dimensions. Note that $\Omega^{\underline \mu}{}_{\underline \rho}$ and $\Omega^{\underline m}{}_{\underline p}$ are parameters of Lorentz transformations that die off at infinity and thus do not contribute to the true zero modes, which correspond to the broken generators of the global asymptotic symmetry group. The zero-modes of the 5-brane in $D=9$ are thus properly balanced between the bose and fermi sectors. With respect to the unbroken $N=1$, $D=6$ supersymmetry, they form a hypermultiplet, in the form where the four scalars occur as an $SU(2)$ triplet plus a singlet. The spinor zero-modes form an $SU(2)$-Majorana doublet. Note that while the zero-modes form a supermultiplet under the unbroken supersymmetry in $d=6$, as they must, not all of the scalars correspond to translational zero-modes. In the reduction from 9 to 6 dimensions, only three translational modes occur, leaving one more to arise from a different broken gauge symmetry. In the present case, this extra scalar mode arises from the broken gauge symmetry of the $A_{\sst M}$ potential, {\it i.e.}\ the $\Lambda$ zero mode in (\ref{zmodeparms}).\footnote{Note also that giving an expectation value to an Abelian field strength can cause symmetry breaking in the present context, unlike in an ordinary Yang-Mills context, because of the structure of the linked gauge symmetry involving $A_{\sst M}$ and $B_{\sst{MN}}$.} The other $p$-brane solutions discussed in this paper all leave half the original $D$-dimensional supersymmetry unbroken, and form appropriate supermultiplets of the unbroken supersymmetry. As another example, consider the string solution in $N=1$, $D=5$ supergravity that we discussed in section 5. Of the original 8 real components of the supersymmetry transformation, 4 are unbroken by the solution and 4 are broken, giving rise to 4 fermionic zero-mode {\it fields}. Of the bosonic zero-modes, there are obviously 3 corresponding to the broken translations. One more scalar zero-mode arises from the broken gauge symmetry of the 2-form field strength. In order to organise these into a supermultiplet of the unbroken $d=2$ supersymmetry, one needs to recall one of the characteristic features of $d=2$ supersymmetry. As one may see from eqn (\ref{stringeps}), the surviving $d=2$ supersymmetry is {\it holomorphic}; it is in fact a (4,0) supersymmetry. This supersymmetry relates the 4 fermionic zero-mode fields to 4 holomorphic bosonic modes. The usual style of counting zero-modes in $d\ge3$, in which the count of fermionic zero-modes is taken to be half of the number of fermionic fields, is not particularly convenient in $d=2$. In $d=2$, the bosons also need to be split into holomorphic and antiholomorphic parts. Although our solitonic string solution clearly will have both holomorphic and antiholomorphic components of the bosonic zero-modes propagating according to the worldsheet equations of motion, only one of these sectors becomes paired with the fermionic zero modes in the (4,0) supermultiplet. The other sector remains unpaired as a set of supersymmetric singlets. \section{Discussion} In this paper, we have searched for supersymmetric $p$-brane solutions of supergravity theories in diverse dimensions. As these solutions occur in families related by dimensional reduction, we have concentrated on the maximal, or stainless, solution in each family. In addition to the previously-known examples, we have found a number of new solutions in $5\le D\le 9$ dimensions. (The lower dimensional bound arises here because we have restricted our attention to solutions of supergravity theories, and have not considered $p$-branes in theories with rigid supersymmetry.) Our new stainless solutions cannot oxidize isotropically to brane solutions in higher dimensions. Put another way, this means that these new solutions cannot simply be viewed as dimensional reductions of previously-known brane solutions in $D=11$ or $D=10$ dimensions. A question that we have not addressed so far concerns the world brane actions that should describe the zero-mode fluctuations around the static, isotropic solutions considered here. From supersymmetry, one has detailed knowledge of the supermultiplet structure of the zero modes that would appear in a gauge-fixed world brane action. In general, one expects that such gauge-fixed actions should be extendable to spacetime supersymmetric and Lorentz invariant actions by adding the appropriate additional unphysical degrees of freedom and their associated local world volume gauge symmetries. For the four classic sequences of super $p$-branes, these actions are generalisations of the $D=10$ superstring \cite{gsch,pol} or $D=11$ supermembrane \cite{berg} actions. Very little is known about the structure of the covariant world volume actions for any of the other $p$-brane cases; this subject remains an important open problem. In the absence of detailed knowledge of the world volume action, some information can be extracted if one assumes that the bosonic sector of the action takes the general form of a Nambu-Goto action coupled to the spacetime metric, dilaton, and a $d$-form gauge potential, with the non-polynomiality removed by the introduction of a world-volume metric $\gamma_{ij}$ as an auxiliary field \cite{bhd}: \begin{eqnarray} I_{\rm brane} &=& \int d^d\xi \Big(-\ft12 \sqrt{-\gamma} \gamma^{ij}\, \partial_i X^{\sst M} \partial_j X^{\sst N}\, g_{\sst{MN}}\, e^{b\phi} +\fft{d-2}{2} \sqrt{-\gamma} \nonumber\\ &&- \fft{1}{d!}\, \epsilon^{i_1\cdots i_d}\, \partial_{i_1} X^{{\sst M}_1}\cdots \partial_{i_d} X^{{\sst M}_d}\, A_{{\sst M}_1\cdots {\sst M}_d}\Big) \ .\label{bract} \end{eqnarray} The exponent $b$ of the dilaton coupling in the first term is {\it a priori} unknown. One way of selecting a value for it in a number of cases is by resorting to an argument based on a scaling symmetry of the pure supergravity theory \cite{dkl}. Under the constant rescalings \begin{equation} e^\phi\longrightarrow \lambda^\alpha\, e^\phi\ ,\qquad g_{\sst{MN}}\longrightarrow \lambda^{2\beta}\, g_{\sst{MN}}\ ,\qquad A_{\sst{M}_1\cdots \sst{M}_d}\longrightarrow \lambda^\gamma\, A_{\sst{M}_1\cdots \sst{M}_d}\ ,\label{scale} \end{equation} the action given by (\ref{boslag}) scales by an overall factor $\lambda^{\beta(D-2)}$, provided that we choose $\gamma=\beta d + a \alpha$. Requiring that $I_{\rm brane}$ scale in the same way, one finds that the parameters must be related by \begin{equation} \alpha=\fft{d{\tilde d}}{a(D-2)}\ ,\qquad \beta=\fft{d}{D-2}\, \qquad \gamma=d\ ,\qquad b=\fft{a}{d}\ .\label{solpar} \end{equation} Substituting the ans\"atze (\ref{metrform}) and (\ref{eleans}) into (\ref{bract}), one finds that the branewave equation following from (\ref{bract}) implies \cite{dkl} \begin{equation} (e^C)'= (e^{dA + a\phi/2})'\ .\label{branewave} \end{equation} Comparing with the solution given by (\ref{solution1}) and (\ref{csol}), one finds that $a$ must satisfy \begin{equation} a^2=4 -\fft{2d{\tilde d}}{D-2}\ .\label{dudelta} \end{equation} Thus requiring that the elementary brane solution should also satisfy the branewave equation, with the parameter $b$ in (\ref{bract}) determined by requiring the above scaling symmetry, the value of $a$ appearing in (\ref{boslag}) is in all cases given by (\ref{avalue}) with $\Delta=4$. All of the antisymmetric tensors in $D=11$ and $D=10$ supergravities have $\Delta=4$, as we have seen. Thus the elementary $p$-brane solutions associated to these antisymmetric tensors accord with the above discussion. However, as we have seen, there are other values of $\Delta$ that also occur in lower dimensional supergravity theories. Elementary $p$-brane solutions in such cases cannot have zero modes that are described by the action (\ref{bract}) with the choice of parameter $b$ dictated by the scaling symmetry. An example is provided by the self-dual string in $D=6$, for which the self-dual 3-form has $\Delta=2$. In this case, since there is no action for $N=1$, $D=6$ supergravity, one has to implement the scaling argument at the level of the equations of motion. The supergravity equations of motion themselves (\ref{d6eqmo}) are invariant under the scaling $g_{\sst{MN}}\longrightarrow \lambda^{2\beta} g_{\sst{MN}}$ and $G_{\sst{MNP}}\longrightarrow \lambda^{2\beta} G_{\sst{MNP}}$, for arbitrary $\beta$. However, the energy-momentum tensor $T_{\sst{MN}}(x)= -\int d^2\xi \sqrt{-\gamma}\gamma^{ij} \partial_i X^{\sst P} \partial_j X^{\sst Q}\, g_{\sst{MP}} g_{\sst{NQ}}\delta^6(x-X)/\sqrt{-g}$ scales with a factor $\lambda^{-2\beta}$. Thus the coupled supergravity-string equations break the scaling symmetry. In fact all of the new stainless elementary $p$-branes, which have $\Delta=2$ or $\Delta=\ft43$, exhibit a similar breaking of the scaling symmetry in their couplings. The ultimate significance of the scaling symmetry used in the above arguments remains unclear to us. Are elementary $p$-brane solutions that respect the scaling symmetry in their couplings more fundamental than others? We do not have an answer to this question at present. We would remark, however, that couplings that break scaling symmetries present in pure uncoupled gravity and supergravity theories are not at all uncommon. Take, for example, a massless charged particle coupled to Einstein-Maxwell theory in $D=4$. The Einstein-Maxwell action $I=\int d^4x (e R -\ft14 e F^2)$ in the absence of the particle coupling scales as $I\longrightarrow \rho^2 I$ under $g_{\mu\nu}\longrightarrow\rho^2 g_{\mu\nu}$, $A_\mu\longrightarrow \rho A_\mu$. But once the particle is coupled in, {\it via} the standard worldline-reparametrisation invariant action $\int d\tau (e^{-1}\dot X^\mu\dot X^\nu g_{\mu\nu} + A_\mu\dot X^\mu)$, the scaling symmetry is broken by the electromagnetic coupling $A_\mu\dot X^\mu$. Another reason why the status of the scaling symmetries remains in doubt comes from quantisation. The supergravity theories arising as low-energy effective field theories of superstring theories have field equations determined {\it via} the beta functions from the requirement that the string's conformal invariance be preserved. The leading terms of these effective field equations reproduce standard supergravity field equations, but there are also an infinite series of quantum corrections, all of which break the scaling symmetries. Nonetheless, it is intriguing that a purely bosonic discussion based upon the coupling of Nambu-Goto-type actions and preservation of scaling symmetries fixes dilaton couplings to antisymmetric tensor gauge fields in a way that agrees with many of the couplings actually found in supergravity theories ({\it i.e.}\ the $\Delta=4$ couplings). In concluding, we shall summarise the results that we have obtained in this paper in a revised brane-scan, in which we plot only the stainless members of each $p$-brane family. In accordance with our discussion at the end of section 2.1, we may, without loss of generality, consider only the versions of the various supergravity theories where all the antisymmetric tensors have degrees $n\le D/2$, since no further inequivalent elementary or solitonic brane solutions arise from dualised versions of the theories. Accordingly, in the stainless brane-scan, we denote solutions of the $n\le D/2$ theories that are elementary by open circles, solutions that are solitonic by solid circles, and self-dual solutions by cross-hatched circles. The dashed lines extending diagonally downwards from the various points on the brane scan indicate that each stainless solution gives rise to its own set of dimensionally-reduced descendants. \newpage \begin{center} {\bf \Large Brane Scan of Stainless Supergravity Solutions} \vskip1cm \epsfbox{bscan.eps} \bigskip\bigskip (For supergravity theories in their $n\le D/2$ versions) \end{center} \section*{Acknowledgements} H.L., C.N.P. and K.S.S. thank SISSA, Trieste, and E.S. thanks ICTP, Trieste, for hospitality during the course of this work.
\section{Introduction} Classical gauge field theories may be defined by a set of fields ${\cal A}$, subject to a set of constraints ${\cal B}$, which, in turn, generate gauge transformations. The quantization of such theories is not a unique procedure. Indeed, already in the two best-established quantization methods very different technical setups are chosen. On the one hand, one has the canonical operator formalism, originating with Heisenberg and Pauli \cite{Heisenberg}, and now well-adapted to handle non-abelian gauge theories \cite{Kugo}, whereas on the other hand Feynman's path integral formalism \cite{Feynman} allows the quantization of such theories through the Faddeev-Popov procedure \cite{Faddeev}. Both methods lead to identical perturbative expansions, but even at a mathematically heuristic level their possible equivalence is only known in perturbation theory. It is certainly of general interest to have as many conceptually and mathematically different quantization schemes as possible, and to examine the particular features of each of them. The hope of obtaining some hints on how to quantize gravity may provide further motivation for investigating new quantization schemes. Especially, the modern formulation of classical mechanics in terms of symplectic manifolds and Poisson algebras (see e.g. \cite{MaRa}) has suggested more refined quantization procedures, such as geometric quantization \cite{Woodhouse}, and strict deformation quantization \cite{Rie91,NPLJGP}. A particular feature of classical gauge theories that should somehow be reflected in the quantization method is that the physical (reduced) phase space may be written as a so-called Marsden-Weinstein quotient \cite{Arms,BFS}. It was shown in \cite{Landsman} that this classical reduction procedure has a satisfactory quantum analogue in a procedure from operator algebra theory known as Rieffel induction \cite{Rieffel}. The way we apply this technique is mainly operator-theoretic, but a certain aspect of the path integral formalism, viz.\ the integration over the gauge group, will play a r\^{o}le as well. This work discusses certain features of the Rieffel induction procedure, as applied to the quantization of constrained systems, which provides a conceptually and technically new method for the quantization of certain gauge field theories. The method in question has already been successfully applied to certain finite-dimensional constrained systems \cite{Landsman}, as well as to free quantum electrodynamics \cite{Landswied,UAW}. The present work draws on these results. Its aim is two-fold. Firstly, we would like to present the strategy of this new quantization method in a form accessible to a wider scientific community. Therefore, in Chapter 2, we briefly review the main line of argument, leading to the quantization proposal. To keep our presentation reasonably short, we refer for some of the technicalities to the aforementioned papers. Subsequently, in Chapter 3 we apply the new quantization scheme to the St\"uckelberg-Kibble model. This toy model has often been used in the investigation of the Higgs mechanism and of spontaneous symmetry breaking, see e.g.\ \cite{Morchio}. Here, we have chosen it since it already shows many of the typical complications of spontaneously broken gauge theories without the need to restrict oneself to a perturbative discussion. As we shall demonstrate explicitly for this model, the Rieffel induction procedure provides a scheme for the construction of the physical state space of a constrained quantum theory, starting from a larger (unphysical) state space on which the unconstrained theory is defined. Our discussion will focus on the particular properties of this Rieffel-induced physical Hilbert space ${\cal H}_{phys}$. Especially, we find that ${\cal H}_{phys}$ carries a trivial representation of the gauge group and a massive representation of the Poincar\'e group. Also, the positive spectrum condition turns out to be satisfied. As an important by-product, we are able to trace back how ``would-be Goldstone bosons rearrange to a massive, longitudinal component'' in a theory exhibiting the Higgs mechanism. The context of our work is modern symplectic geometry and reduction theory on the classical side, and algebraic quantum field theory on the quantum side. We only use the `soft' side of these theories. Good recent introductions are \cite{MaRa,Haag,BW}, respectively. \section{The quantization of gauge theories with Rieffel induction} After presenting schematically the strategy which leads to Rieffel induction in the quantization of theories with constraints, the remainder of this section briefly specifies some notational and technical prerequisites. \subsection{Quantization of Marsden-Weinstein reduction} The general symplectic reduction procedure, which is quantized by Rieffel induction in its full generality, is described in \cite{Landsman}. Here we are merely concerned with a special case, viz.\ Marsden-Weinstein reduction at the zero level of the moment map, cf.\ \cite{MaRa,MW}. To introduce our notation, let us consider free classical electrodynamics. For the functional-analytic and other details which are suppressed in what follows, we refer the interested reader to \cite{Landswied}. We start with the space $M$ of four-component real-valued weak solutions $A_{\mu}$ of the wave equation whose Fourier-transformed Cauchy-data lie in $ L^2({\Bbb R}^3)\otimes {\Bbb C}^4 $. That is, $M = {\lbrace{ A_{\mu} | {\Box}A_{\mu} = 0 }\rbrace}$. The imaginary part \begin{equation} B(A,A') = 2{\rm Im}(A,A')_M = -i \int {d^3{\bf p}\over (2\pi)^3} {\lbrack{A^{\mu}({\bf p})\overline{A'}_{\mu}({\bf p}) - \overline{A}^{\mu}({\bf p}) {A'}_{\mu}({\bf p})}\rbrack} \ll{sympl-form} \end{equation} of the indefinite covariant scalar product $(*,*)_M$ turns $M$ into a symplectic space $(M,B)$, which is the phase space of the unconstrained classical system. The set of constraints is given by the gauge group $G$, which acts on $M$ via $A_{\mu} \to A_{\mu} + {\partial}_{\mu}g$, where \begin{equation} G = {\lbrace{ g\in {\cal S}'({\Bbb R}^4)\mid {\Box}g=0; dg \in M }\rbrace}. \end{equation} Here, the space of distributions ${\cal S}'({\Bbb R}^4)$ is the dual of the usual Schwartz space of rapidly decreasing test functions. In the present example, the reduced phase space $(M_c,B_c)$ of the corresponding constrained system may be obtained by a so-called Marsden-Weinstein reduction \cite{MW}. This involves the moment map $J$ from $M$ into the dual of the Lie algebra of $G$. As $G$ is a vector space, we may identify it with its Lie algebra, so we simply write $J_g(A)$ for the value of $J(A)$ on $g\in G$. Explicitly, the moment map turns out to be $J_g(A) = {\rm Im}({\partial}g,A)_M$, cf. \cite{Landswied}. The preimage of its zero level is \begin{equation} J^{-1}(0) = {\lbrace{ A_{\mu}\in M \mid {\partial}_{\mu}A^{\mu} = 0}\rbrace}. \end{equation} Then, $M_c$ is given by the Marsden-Weinstein quotient \begin{equation} M_c = J^{-1}(0) / G, \end{equation} and $B_c$ inherits its structure from $B$. It is easy to see that $(M_c,B_c)$ defined this way indeed describes the physical degrees of freedom of free electrodynamics: picking $J^{-1}(0)$ fixes the gauge (thus imposing the Gauss law constraint, which on elements of $M$ becomes the Lorentz gauge condition), and quotienting by $G$ removes the gauge degeneracy of the symplectic form $B$ with respect to the action of $G$ on $J^{-1}(0)$. In principle, there are two possibilities to quantize a reduced phase space $(M_c,B_c)$. Either, we directly quantize the Marsden-Weinstein reduced (i.e. constrained) classical system $(M_c,B_c)$, or we quantize the unconstrained classical system $(M,B)$ together with the set of constraints. In the latter case, a scheme has to be found which imposes constraints on the unconstrained quantized theory, thereby providing a quantum analogue of the classical Marsden-Weinstein reduction. Examples of such schemes are the Dirac or the BRST method. According to the proposal of \cite{Landsman}, the so-called Rieffel induction procedure of operator algebra theory \cite{Rieffel} (which we explain below) provides a rival scheme, which in all examples studied so far works as well as, or better than the methods mentioned above. More precisely, let us consider schematically a quantization prescription $Q_{\hbar}$ which relates the symplectic space $(M,B)$ (or rather the Poisson algebra of functions on it) to some algebra of field operators on a Hilbert space ${\cal A}$, $G$ to some algebra ${\cal B}$ generated by $G$, and $(M_c,B_c)$ to some (a priori unknown) algebra of observables (in the sense of gauge-invariant operators) ${\cal A}_{obs}$. Then, according to our quantization proposal, the following diagram commutes: \begin{equation} \begin{array}{ccc} (M,B) ; G & \stackrel{Q_{\hbar}}{\longrightarrow}& {\cal A}; {\cal B}\\ \vcenter{ \llap{$\scriptstyle{\rm Marsden-Weinstein\,\, Reduction}$}} \Big\downarrow& &\Big\downarrow\vcenter{ \rlap{$\scriptstyle{\rm Rieffel\,\, Induction}$}}\\ (M_c,B_c) & \stackrel{Q_{\hbar}}{\longrightarrow}& {\cal A}_{obs} \end{array} \ll{diag1} \end{equation} Our program in this paper is to specify the entries of this diagram for the St\"uckelberg-Kibble model. To this end, we briefly recall how, for a linear field theory, a symplectic space $(M,B)$ can be related to a field algebra ${\cal A}$ of canonical commutation relations, and we explain how Rieffel induction allows one to construct new Hilbert spaces for quantum field theories, thereby eventually specifying ${\cal A}_{obs}$. \subsection{Weyl algebras of canonical commutation relations} The general theory behind this subsection is explained in great detail and rigour in, e.g., \cite{BW}, and the application to electromagnetism is from \cite{CGH}. We merely mention some of the main points. For ${\phi}, {\phi}' \in M$, the operators $W({\phi})$, $W({\phi}')$, satisfying the Weyl form of the canonical commutation relation (CCR) \begin{equation} W({\phi})W({\phi}') = W({\phi}+{\phi}') e^{{-i\over 2}B({\phi},{\phi}')}, \ll{ccr-qed-weyl} \end{equation} specify a field algebra with $C^*$-structure which we denote by ${\cal A}(M,B)$. In most cases, one is primarily interested in the properties of the operator vector potential $A_{\mu}$, for which we use the same notation as for its classical counterpart, as no confusion will arise. The $A_{\mu}$ satisfy the canonical commutation relations \begin{equation} [A_{\mu}(x), A_{\nu}(y)] = -ig_{\mu\nu} D(x-y),\ll{ccr-qed-vect} \end{equation} where $D$ denotes the commutator function satisfying $\Box{D}=0$, with initial conditions $D({\bf x},0)=0$, ${\partial\over{\partial t}} D({\bf x},t){\mid}_{t=0} = -{\delta}^{(3)}({\bf x})$. To see the connection between (\ref{ccr-qed-weyl}) and (\ref{ccr-qed-vect}), we consider the vector potential $A(f) = \int d^4x A_{\mu}(x)f^{\mu}(x)$, smeared with real test functions $f$. Now, (\ref{ccr-qed-vect}) reads $[A(f),A(g)]=i\sigma(f,g)$, where $\sigma(f,g) = -\int d^4x d^4y D(x-y)f^{\mu}(x)g_{\mu}(y)$. Formally, this allows for the introduction of the operators $U(f)=e^{[iA(f)]}$ which according to the Baker-Campbell-Haussdorff formula satisfy the Weyl form of the canonical commutation relations $U(f)U(g) = U(f+g) e^{[-{i\over 2}\sigma(f,g)]}$. Here, however, $U(f)$ and $U(f')$ have the same commutation relations as long as $\int d^4x D(x-y) (f^{\mu}(x) - {f'}^{\mu}(x)) = 0$ for almost all $y$. To remove this degeneracy and to obtain a one-to-one correspondence between Weyl operators and test functions, one uses the map $f\to\phi$, defined by the convolution ${\phi}_{\mu} = D*f_{\mu}$. Then, the space $M$ of solutions of the wave equation $\Box{\phi}_{\mu}=0$, $ {\phi}_{\mu}({\bf x},t) = {1\over {{(2\pi)}^3}} \int {d^3{\bf k}\over 2k_0} [{\phi}_{\mu}({\bf k})e^{-ikx}$ $+ \overline{{\phi}_{\mu}}({\bf k})e^{ikx}]$, is\footnote{ Our notation does not distinguish between functions $\phi$ and their Fourier transforms, since no confusion should arise.} \begin{equation} M = \overline{\lbrace{ \phi = D*f}\rbrace} = L^2({{\Bbb R}}^3)\otimes{{{\Bbb C}}\,}^4. \ll{M-qed} \end{equation} Now, the operators $W(\phi) = U(f)$, $\phi\in M$ satisfy (\ref{ccr-qed-weyl}) with symplectic form $B$ induced by $\sigma$ and given in (\ref{sympl-form}). Having established the connection between Weyl operators and vector potentials, we can introduce formal annihilation and creation operators $a_{\mu}$, $a_{\mu}^*$. E.g. for the free electromagnetic field, $ A_{\mu}(x) = \int {d^3{\bf k}\over (2\pi)^32k_0}\, {\lbrack{e^{-ikx} a_{\mu}({\bf k}) + e^{ikx} a_{\mu}^*({\bf k})}\rbrack}{\mid}_{k_0={\bf k}}$, \begin{equation} iA(f) = \int {d^3{\bf k}\over {(2\pi)}^32k_0} {\lbrack{ a_{\mu}({\bf k})\overline{{\phi}_{\mu}}({\bf k}) - a_{\mu}^*({\bf k}) {\phi}_{\mu}({\bf k})}\rbrack} =: a_{\mu}({\phi}^{\mu}) - a_{\mu}({\phi}^{\mu})^*. \ll{crean} \end{equation} Clearly, in terms of the annihilation and creation operators, the Weyl operators read $ W({\phi}^{\mu}) = \exp{\lbrack{a_{\mu}({\phi}^{\mu}) - a_{\mu}({\phi}^{\mu})^*}\rbrack}$, where ${\lbrack{a_{\mu}({\phi}^{\mu}), a_{\nu}({{\phi}'}^{\nu})^*}\rbrack} = ({\phi}',{\phi})_M$, $(.,.)_M$ denoting the indefinite Minkowski inner product. Heuristically, one has \begin{equation} {d\over d\lambda}W(\lambda{\phi})|_{\lambda=0} = iA(f).\ll{deriv-W} \end{equation} It is well-known that this derivative does not exist in the operator norm but with respect to regular representations only, and thereby the $a_{\mu}$, $a_{\mu}^{*}$ only exist in such representations, too. Nevertheless, in what follows we shall adopt the formal expressions (\ref{crean}) and (\ref{deriv-W}), even when no explicit reference to a particular representation is made. As a final preparatory step, we point out that subalgebras of ${\cal A}(M,B)$ can be specified by selecting subspaces of $M$. In particular, for free QED, \begin{eqnarray} N &=& {\lbrace{ {\phi}_{\mu}\in M | k^{\mu}{\phi}_{\mu}({\bf k}) = 0} \rbrace} = {\lbrace{ {\phi}_{\mu}\in M | {\partial}^{\mu} {\phi}_{\mu}(x) = 0}\rbrace}, \nonumber \\ T &=& {\lbrace{ {\phi}_{\mu}\in M | {\phi}_{\mu}({\bf k}) = ik^{\mu}g({\bf k})}\rbrace} = {\lbrace{ {\phi}_{\mu}\in M | {\phi}_{\mu}(x) = {\partial}^{\mu}g(x), \Box g(x) = 0}\rbrace} \nonumber \\ \ll{subspace} \end{eqnarray} define subalgebras ${\cal A}(N,B)$, ${\cal A}(T,B)$ of ${\cal A}(M,B)$. Note that $T\subset N$, so that ${\cal A}(T,B)$ $\subset {\cal A}(N,B)$. These subalgebras are Poincar\'e-invariant, as may be seen by recalling that the action of elements $(\Lambda, a)$ of the Poincar\'e group ${\cal P}$ on ${\cal A}(M,B)$ is defined via the algebraic automorphism ${\alpha}_{(\Lambda, a)}$, \begin{equation} {\alpha}_{(\Lambda, a)}(W({\phi}^{\mu})) = W({\gamma}_{(\Lambda, a)}({\phi}^{\mu})) \qquad\mbox{with}\qquad ({\gamma}_{(\Lambda, a)}({\phi}^{\mu}))(x) = {\Lambda}^{\mu}_{\nu}{\phi}^{\nu}({\Lambda}^{-1}(x-a)).\ll{Poincare} \end{equation} \subsection{Rieffel induction} This subsection gives a quick `review by example' of some parts of the theory developed in \cite{Landsman} and \cite{Landswied}. In physics, induction methods are mainly known from Wigner's classification and construction of all irreducible unitary representations of the Poincar\'e group $P$. In general, the method of induced representations of (locally compact) groups allows one to construct a representation of the complete group from a representation of a subgroup, cf.\ e.g.\ \cite{Barut}. Also, in the theory of operator algebras (particularly $C^*$-algebras) a method exists for constructing a representation of an algebra, given a representation of some other algebra \cite{Rieffel}. The latter is not necessarily a subalgebra of the former; instead, the two algebras need to be connected by a bimodule with certain additional properties. Whatever the technical details, the main idea is that the representation one induces from should be straightforward, and yet capable of producing an appropriate representation of the algebra one is really interested in. This idea will be fully realized in our context, for the second algebra will be the algebra generated by the gauge group, and the representation induced from is the trivial one. With a suitable choice of bimodule, the induced representation of the algebra of observables comes out to be the vacuum representation on a Fock space of physical photon states. To facilitate our presentation, we proceed by example, abstracting general features afterwards. For free QED, in the diagram (\ref{diag1}) we choose the field algebra ${\cal A} = {\cal A}(M,B)$ and the `algebra of constraints' ${\cal B} = {\cal A}(T,B)$ (cf.\ the previous subsection), where the choice of ${\cal B}$ is motivated by observing that the gauge group $G$ equals $T$, cf. (\ref{subspace}).\footnote{ For simplicity, we here ignore some mathematical difficulties in defining algebras ${\cal B}$ for groups $G$ which are not locally compact. This greatly simplifies our presentation. For more details, we refer to \cite{Landswied,UAW}.} Also, we introduce the `algebra of weak observables' ${\cal A}_c := {\cal A}(N,B)$, which is the largest subalgebra of ${\cal A}={\cal A}(M,B)$ commuting with ${\cal B} = {\cal A}(T,B)$. The Rieffel induction procedure will produce a representation of ${\cal A}_c$ induced from a representation of $\cal B$. To this end, we need a bimodule for ${\cal A}_c$ and $\cal B$, that is, a linear space on which ${\cal A}_c$ acts from the left, and $\cal B$ acts from the right (that is, in an anti-representation), so that these two actions commute. In the case at hand, ${\cal A}_c$ and ${\cal B}$, which is abelian, are each other's commutant in the field algebra $\cal A$, so that a representation of $\cal A$ on a Hilbert space $\cal H$ automatically defines such a bimodule. Finally, we need a representation of $\cal B$ to induce from. This is the trivial one, defined on the Hilbert space ${\cal H}_{tr}={\Bbb C}$. Schematically, \begin{equation} {\cal A}_c \longrightarrow {\cal H} \longleftarrow {\cal B} \longrightarrow {\cal H}_{tr}. \ll{diag2} \end{equation} The restriction of the action $\pi$ on $\H$ of $\cal A$ to its subalgebra $\cal B$ defines a representation $U$ of the gauge group, that is, one has $U(\phi)=\pi(W(\phi))$. As will be discussed in more detail below, this setup allows the construction of a positive semidefinite sesquilinear form $(.,.)_0$ on $L\otimes {\cal H}_{tr}$, where $L$ is a suitable dense subspace of $L$. In the present case, this form is given by \begin{equation} (\psi\otimes v,\varphi \otimes w)_0 = v\overline{w} \int_G [{\cal D}\phi] (U(\phi) \psi,\varphi). \ll{0-prod} \end{equation} Here $[{\cal D}\phi]$ denotes the non-existent `Lebesgue' measure on the gauge group $G$. The point is, however, that this flat `measure' combines with a factor in the integrand to define a mathematically well-defined path integral (cylindrical) measure on $G$ \cite{Landswied}. Furthermore, $\psi, \varphi $ are in $ {\cal H}$, $v,w \in {\cal H}_{tr} = {{\Bbb C}}$, and $(.,.)$ is the inner product on $\H$. Irrespective of the explicit form of $(.,.)_0$, the induced physical Hilbert space is then defined as the completion of the quotient of $L\otimes {\cal H}_{tr}$ by the null space of $(.,.)_0$, i.e., \begin{equation} {\cal H}_{phys} = \overline{(L\otimes {\cal H}_{tr}) / {\cal N}}, \ll{quotient} \end{equation} where ${\cal N} \subset L\otimes {\cal H}_{tr}$ is the subset of vectors with vanishing $(.,.)_0$ norm. The collection of vectors in ${\cal H}_{phys}$ of the form $\psi\tilde{\otimes}v$, defined as the image of $\psi\otimes v\in L\otimes {\cal H}_{tr}$ under the quotient projection from $L\otimes {\cal H}_{tr}$ to ${\cal H}_{phys}$, are clearly dense in ${\cal H}_{phys}$. The action of elements $A$ of ${\cal A}_c$ on ${\cal H}_{phys}$ is then given on this dense set by ${\pi}_{phys}(A)\psi\tilde{\otimes}v = ({\pi}(A)\psi)\tilde{\otimes}v$. Under appropriate continuity conditions \cite{Rieffel,Landswied} this action may be extended to all of ${\cal H}_{tr}$. The reader should note that ${\cal H}_{phys}$ satisfies an essential requirement of a non-degenerate physical Hilbert space: the gauge degeneracy of elements of ${\cal A}_c$ is removed in ${\pi}_{phys}({\cal A}_c)$. To see this, choose an arbitrary element $W(\phi) \in {\cal A}_c$. From equation (\ref{0-prod}), it is obvious that for ${\phi}_t \in T$ (which, we recall, coincides with the gauge group $G$), ${\pi}_{phys}(W(\phi))\psi\tilde{\otimes}v = {\pi}_{phys}(W(\phi + {\phi}_t))\psi\tilde{\otimes}v$ for all vectors $\psi\tilde{\otimes}v \in {\cal H}_{phys}$. Hence, ${\pi}_{phys}(W(\phi)) = {\pi}_{phys}(W(\phi + {\phi}_t))$. This removal of the gauge degeneracy of ${\cal A}_c$ is independent of the choice of ${\cal H}$, and hence we indentify ${\pi}_{phys}({\cal A}_c)$ with the representation-independent algebra of observables ${\cal A}_{obs}$, cf.\ (\ref{diag1}). Let us now turn to the abstract setting which has led to the $(.,.)_0$-inner product (\ref{0-prod}). As stated, the aim of the Rieffel induction procedure is to obtain a representation ${\pi}_{phys}$ of ${\cal A}_c$ induced from a representation of ${\cal B}$ on some Hilbert space ${\cal H}_{\chi}$. Our example, and all similar examples involving gauge theories, have the special feature that ${\cal H}_{\chi}={\cal H}_{tr}={{\Bbb C}}$, that is, one induces from the trivial representation of the gauge group. This will imply that the algebra of constraints ${\cal B}$ is represented trivially on the induced space ${\cal H}_{phys}$. Technically, the construction of ${\pi}_{phys}$ proceeds according to the following three step method. \begin{enumerate} \item Given a bimodule $L$ for ${\cal A}_c$ and $\cal B$, a ${\cal B}$-valued scalar product ${\langle{.,.}\rangle}_{\cal B}$ has to be found on $L$, that is, for $\psi, \varphi \in L \subset {\cal H}$, ${\langle{\psi,\varphi}\rangle}_{\cal B} \in {\cal B}$,\footnote{ Mathematically, ${\langle{.,.}\rangle}_{\cal B}$ is a so-called rigging map which has to satisfy the following conditions for all $\psi$, $\varphi\in L$ \cite{Rieffel}: \begin{enumerate} \item ${\langle{\lambda\psi,\mu\varphi}\rangle}_{\cal B} = \overline{\lambda}\mu{\langle{\psi,\varphi}\rangle}_{\cal B}$ $\hbox{ }$ for all ${\lambda},\mu\in{{\Bbb C}}$; \item ${\langle{\psi,\varphi}\rangle}_{\cal B}^* = {\langle{\varphi,\psi}\rangle}_{\cal B}$ (where the $\mbox{}^*$ denotes the hermitian conjugate in $\cal B$); \item ${\langle{\psi,\varphi B}\rangle}_{\cal B} = {\langle{\psi,\varphi}\rangle}_{\cal B}B$ $\hbox{ }$ for all $B\in {\cal B}$ (on the left-hand side, $B$ acts in the given right-representation on the bimodule $L$, whereas on the right-hand side $B$ acts by multiplication in the algebra $\cal B$); \item ${\langle{A\psi,\varphi}\rangle}_{\cal B} = {\langle{\psi,A^*\varphi}\rangle}_{\cal B}$ $\hbox{ }$ for all $A\in {\cal A}$. \item ${\langle{A\psi,A\psi}\rangle} \leq {\parallel{A}\parallel}^2 {\langle{\psi,\psi}\rangle}$ $\hbox{ }$ for all $\psi \in L$, $A\in {\cal A}$. \end{enumerate}} \item Given such an operator-valued scalar product, the tensor product $L\otimes {\cal H}_{\chi}$ is equipped with a sesquilinear form $(.,.)_0$, \begin{equation} (\psi\otimes v,\varphi\otimes w)_0 := ({\pi}_{\chi}({\langle{\varphi, \psi}\rangle}_{\cal B})v,w)_{\chi}.\ll{rigged} \end{equation} Crucially, this form is positive-semidefinite if the postivity condition \\ $\pi_{\chi} ({\langle{\psi,\psi}\rangle}_{\cal B})\geq 0$ for all $\psi\in L$ is satisfied, which is the case in all our examples. \item The subspace ${\cal N} \subset L\otimes {\cal H}_{tr}$ of vectors with vanishing $(.,.)_0$-norm is determined and the physical Hilbert space is defined as in (\ref{quotient}). \end{enumerate} The most difficult part of this procedure is to find ${\langle{.,.}\rangle}_{\cal B}$. Here, one is guided by mathematical examples \cite{Landsman}. One may consider e.g. ${\cal B} = C^*(G)$, the $C^*$-group algebra of a locally compact group $G$ (cf.\ \cite{BW}; this is essentially the convolution algebra on the group w.r.t.\ the Haar measure). Then, it can be shown that a rigging map ${\langle{.,.}\rangle}_{\cal B}$ is defined as follows: ${\langle{\psi,\varphi}\rangle}_{\cal B}$ has to be some element of $C^*(G)$, i.e., a function on the group, and we prescribe that the value of this function at $g\in G$ is given by ${\langle{\psi,\varphi}\rangle}_{\cal B} (g) = (U(g)\varphi,\psi)$, where $U$ is a continuous unitary representation of $G$ on ${\cal H}$, commuting with ${\pi}({\cal A}_c)$, $x\in G$. Inducing from the trivial representation ${\cal H}_{tr} = {{\Bbb C}}$, one obtains\footnote{ In what follows, we use the shorthand $(\psi,\varphi)_0$ for $(\psi\otimes v,\varphi\otimes w)_0$, since $v,w \in {{\Bbb C}}$ are complex numbers which can be absorbed in the definition of $\psi$ and $\varphi$.} \begin{equation} (\psi,\varphi)_0 = \int_G dx (U(x)\psi,\varphi), \ll{0-prod-G} \end{equation} of which (\ref{0-prod}) is a special case, at least in a heuristic sense. In what follows, we shall take a suitable generalization of (\ref{0-prod-G}) as our starting point, thereby obviating the need for a discussion of the explicit form and a verification of the mathematical properties of ${\langle{.,.}\rangle}_{\cal B}$. In fact, our presentation of the Rieffel induction procedure for quantum field theories has been slightly oversimplified with respect to this point. While the existence of a so-called `rigged' inner product $(.,.)_0$, defined in (\ref{rigged}), is always sufficient for the quantization proposal to apply, it is not always possible to derive it from a mathematically well-defined rigging map ${\langle{.,.}\rangle}_{\cal B}$. We refer to \cite{Landswied} for a discussion of the technical points involved. To sum up: In this chapter, we have seen that Rieffel induction provides a well-defined scheme for the construction of a physical Hilbert space ${\cal H}_{phys}$, on which gauge transformations act trivially. In the corresponding algebra of observables ${\pi}_{phys}({\cal A}_c)$, all gauge degeneracies are removed, i.e., Rieffel induction is a method to impose constraints on quantum field theories. The physical Hilbert space ${\cal H}_{phys}$ is obtained by forming the quotient of a larger Hilbert space $L\otimes {\cal H}_{tr}$ with respect to a null space. This is somehwat reminiscent of the BRST (or, in case of QED, the Gupta-Bleuler) procedure, with the major difference that with Rieffel induction no negative-norm subspace exists, obviating the need to select a physical subspace of $\cal H$. Also, certain functional-analytic problems that appear in the BRST as well as in the Dirac method are absent with our present techniques \cite{Landsman,Landswied}. By definition of the inner product on the physcial Hilbert space ${\cal H}_{phys}$, calculations of correlation functions of operators in ${\cal A}_c$ (as represented on ${\cal H}_{phys}$) may be performed in $L\otimes {\cal H}_{tr}$, \cite{UAW}. \section{Application to the St\"uckelberg-Kibble model} In this chapter, we specify (\ref{diag2}) and (\ref{0-prod}) for the St\"uckelberg-Kibble model, thereby constructing a physical Hilbert space ${\cal H}_{phys}$ for this model. The St\"uckelberg-Kibble model is an abelian Higgs model with the modulus $\eta$ of the scalar field $\phi(x) = \eta(x) e^{\varphi(x)}$ frozen to unity, $\eta(x) =1$. It is given by the Lagrangian \begin{equation} {\cal L}= -{1\over 4} F_{\mu\nu}F^{\mu\nu} - {1\over 2} \left( {\partial}_{\mu}\varphi + eA_{\mu}\right)\left( {\partial}^{\mu}\varphi + eA^{\mu}\right). \end{equation} Despite its linearity, this model has non-trivial features, and has been used as testing ground for investigations of the Higgs mechanism before \cite{Morchio}. Its equations of motion can be written in terms of a gauge-invariant current $j^{\mu} = {\partial}^{\mu}\varphi + e A^{\mu}$, satisfying \begin{equation} \left( \Box + e^2\right) j^{\mu} = 0 \qquad \hbox{;}\qquad {\partial}_{\mu}j^{\mu} = 0. \ll{j-equ} \end{equation} In fact, this is nothing but the Proca equation \cite{Itzykson} of a massive gauge-invariant vector field. To make this model amenable to treatment by symplectic reduction and quantum induction methods, we now make a move that is analogous to rewriting the Maxwell equation for $A_{\mu}$ as a massless Klein-Gordon equation plus a subsidiary Lorentz condition. Thus we pass back to the gauge-dependent fields $A_{\mu}$ and $\varphi$, and choose what is essentially the 't Hooft gauge as the subsidiary condition: \begin{equation} {\partial}_{\mu} A^{\mu} = e\varphi. \ll{thg} \end{equation} With this constraint, the equations of motion read \begin{equation} \left( \Box + e^2\right) A^{\mu} = 0 \qquad \hbox{,} \qquad \left( \Box + e^2\right) \varphi = 0, \ll{g-equ} \end{equation} and the gauge group $G={\lbrace{g| \left( \Box + e^2\right)g = 0}\rbrace}$ acts on $A_{\mu}$, $\varphi$ via \begin{equation} A_{\mu} \to A_{\mu} + {\partial}_{\mu}g \qquad\mbox{ , }\qquad \varphi \to \varphi - eg. \ll{g-trafo} \end{equation} \subsection{Marsden-Weinstein reduction for the St\"uckelberg-Kibble model} A mathematically rigorous treatment of the following material, in the style of \cite{Landswied}, is possible, but we leave the details to the interested reader; instead, readability commands us to give somewhat loose formulations. Our investigation of the St\"uckelberg-Kibble model starts from the symplectic space $(M_{sk},B_{sk})$, defined by \begin{eqnarray} && M_{sk} = {\lbrace{ ( A_{\mu},{\varphi}) \mid A_{\mu}\in L^2({{\Bbb R}}^3)\otimes {{{\Bbb C}}\,}^4, {\varphi}\in L^2({{\Bbb R}}^3); \left( \Box + e^2\right) A_{\mu}= \left( \Box + e^2\right){\varphi}=0 }\rbrace},\nonumber \\ &&B_{sk}( A_{\mu},{\varphi};{A'}_{\mu},{\varphi}') = 2{\rm Im}( A_{\mu},{A'}_{\mu})_M - 2{\rm Im}({\varphi},{\varphi}'). \ll{sk-sympl} \end{eqnarray} The gauge group $G$ acts on this space by the gauge transformation (\ref{g-trafo}). This action is strongly Hamiltonian, and hence, in particular, it is symplectic. We evidently may identify the gauge group with the following subspace of $M_{sk}$ \begin{equation} T_{sk} = {\lbrace{ ( A_{\mu},{\varphi}) \in M_{sk} \mid A_{\mu}= {\partial}_{\mu}g, {\varphi}= -eg; g\in L^2({{\Bbb R}}^3); \left( \Box + e^2\right) g = 0}\rbrace}.\ll{tsk} \end{equation} {}From this, the Marsden-Weinstein reduced space $(M_{c,sk},B_{c,sk})$ is easily calculated. With similar notation as in subsection 2.1, the moment map reads \begin{equation} J_g( A_{\mu},\varphi) = 2{\rm Im}({\partial}_{\mu}g,A_{\mu})_M - 2{\rm Im}(-eg,{\varphi}), \end{equation} which leads to \begin{equation} J^{-1}(0) = \{( A_{\mu},{\varphi}) \in M_{sk} | {\partial}_{\mu} A^{\mu} = e\varphi\}. \ll{jmoo} \end{equation} Hence in view of (\ref{thg}) the Marsden-Weinstein quotient reads \begin{equation} M_{c,sk} = J^{-1}(0)/G = {\lbrace{ j_{\mu} \in L^2({{\Bbb R}}^3)\otimes {{{\Bbb C}}\,}^4| \left( \Box + e^2\right)j_{\mu}=0 ; {\partial}_{\mu}j^{\mu} = 0}\rbrace}. \end{equation} The symplectic form $B_{c,sk}$ on $M_{c,sk}$ inherits its structure from $B_{sk}$, and is given by \begin{equation} B_{c,sk}(j,j') = {2\over e^2} {\rm Im}(j_{\mu}, {{j}'}_{\mu})_M. \end{equation} Clearly, $(M_{c,sk},B_{c,sk})$ is the phase space of a massive vector boson, which indeed represents the physical degrees of freedom of the St\"uckelberg-Kibble model. This completely specifies the left-hand side of the diagram (\ref{diag1}). \subsection{Rieffel induction for the St\"uckelberg-Kibble model} \subsubsection{Construction of the field algebra} Consider the canonical commutation relations of the operator fields $A_{\mu}$ and $\varphi$ (denoted by the same symbol as their classical counterparts): \begin{eqnarray} [{\varphi}(x),{\varphi}(y)] &=& i{\triangle}(x-y), \nonumber \\ {[ A_{\mu}(x), A_{\nu}(y)]} &=& -ig_{\mu\nu}{\triangle}(x-y), \ll{ccrsk} \end{eqnarray} where the commutator function ${\triangle}$ satisfies $(\Box + e^2){\triangle}(x) =0$ with initial conditions ${\triangle}({\bf x},0) =0$, ${{\partial}\over {\partial}t} {\triangle}({\bf x},t){\mid}_{t=0} = -{\delta}^{(3)}({\bf x})$. In analogy with our discussion of free QED, we specify the formal connection between the fields $A_{\mu}$, $\varphi$ and the corresponding Weyl operators, $ W({\phi}_{\mu},{\phi}) = e^{iA_{\mu}(f^{\mu}) + i\varphi(f)},$ where ${\phi}_{\mu} = \triangle * f_{\mu}$, ${\phi} = \triangle * f$. Here, either as a consequence of (\ref{ccrsk}), or imposed axiomatically, the operators $W({\phi}_{\mu},{\phi})$, $W({{\phi}'}_{\mu},{\phi}')$ satisfy the Weyl form of the canonical commutation relations \begin{equation} W({\phi}_{\mu},{\phi})W({\phi}_{\mu}',{\phi}') = W({\phi}_{\mu} + {\phi}_{\mu}',{\phi} + {\phi}') e^{-{i\over 2}B_{sk}({\phi}_{\mu},{\phi};{{\phi}'}_{\mu},{\phi}')}. \end{equation} The field algebra of the model is then defined as the Weyl algebra ${\cal A}(M_{sk},B_{sk})$ generated by the $W$'s subject to these commutation relations (cf.\ \cite{BW}). Now, we want to construct the quantum counterpart of Marsden-Weinstein reduction, i.e., we want to complete the right hand side of the diagram (\ref{diag1}). Therefore, we invoke the quantization prescription for symplectic spaces as discussed in Chapter 2. This leads to the field algebra ${\cal A}\equiv {\cal A}(M_{sk},B_{sk})$ defined by (\ref{sk-sympl}). Also, in analogy with our discussion in Chapter 2, we choose the algebra of constraints ${\cal B}={\cal A}(T_{sk},B_{sk})$; once again, the motivation for this is that it is the ($C^*$) algebra generated by the gauge group. Consequently, the algebra of weak observables, which by definition is the largest subalgebra of ${\cal A}(M_{sk},B_{sk})$ commuting with ${\cal B}(T_{sk},B_{sk})$, is given by ${\cal A}_c={\cal A}(N_{sk},B_{sk})$, where \begin{equation} N_{sk} = {\lbrace{ ({\phi}_{\mu},{\phi}) \mid {\partial}^{\mu}{\phi}_{\mu}= e{\phi} }\rbrace} \subset M_{sk}; \ll{nsk} \end{equation} compare this with (\ref{jmoo}). The subspaces $N_{sk}$ and $T_{sk}\subset N_{sk}$ of $M_{sk}$ are invariant under the action of symplectic transformations ${\gamma}_{\Lambda,a}$ associated with elements $(\Lambda,a)$ of the Poincar\'e group ${\cal P}$, $({\gamma}_{\Lambda,a}({\phi}_{\mu},{\phi}))(x) := ({\Lambda}^{\nu}_{\mu}{\phi}_{\nu},{\phi})({\Lambda}^{-1}(x-a)))$, cf. (\ref{Poincare}). Consequently, the subalgebras ${\cal A}_c$ and ${\cal B}$ are Poincar\'e-invariant. \subsubsection{Representing the algebra of observables} Rieffel induction starts from the input data of diagram (\ref{diag2}). So far, we have determined the algebra of weak observables ${\cal A}_c={\cal A}(N_{sk},B_{sk})$ and the algebra of constraints ${\cal B}={\cal A}(T_{sk},B_{sk})$ of the St\"uckelberg-Kibble model; note that ${\cal B}\subset {\cal A}_c$. What is needed is a representation of these algebras on some subspace $L$ of a Hilbert space ${\cal H}$. In this subsection, we give such a representation on a bosonic Fock space (cf.\ the corresponding procedure for QED in \cite{Landswied}). For simplicity, in a first step we introduce a representation for elements $W({\phi}_{\mu},{\phi}=0) \in {\cal A}(M_{sk},B_{sk})$ only. This will subsequently be generalized to the whole algebra. We start from the canonical commutation relations for the smeared annihilation and creation operators $\hat{a}_{\mu}$, $\hat{a}_{\mu}^*$, \begin{equation} \hat{a}(f)= \hat{a}_{\mu}(f^{\mu}) = \int {d^3{\bf k}\over (2\pi)^32k_0} \lbrack{ \hat{a}_0({\bf k})\overline{f}_0({\bf k}) + \hat{a}_i({\bf k})\overline{f}_i({\bf k})}\rbrack, \end{equation} namely \begin{equation} {\lbrack{ {\hat{a}(f)}, {\hat{a}^*(g)} }\rbrack} = (g,f)_E := \int {d^3{\bf k}\over (2\pi)^32k_0} {g}_{\mu}({\bf k}) {\delta}^{\mu\nu} \overline{f}_{\nu}({\bf k}). \ll{ccr-euclid} \end{equation} For reasons to become clear soon, we have employed the so-called Fermi trick \cite{CGH} which consists in defining the creation and annihilation operators of a vector field such that their commutator is a Euclidean scalar product. Introducing a vacuum state $|0\rangle$ with the property $ \hat{a}(f)|0\rangle=0$ for all $f$, the creation and annihilation operators generate a bosonic Fock space $\H_1$ in the usual way. Mathematically $\H_1$ is, of course, the symmetric Hilbert space \cite{Guichardet} over $L^2({{\Bbb R}}^3)\otimes {{\Bbb C}}^4$. We can now represent the field algebra $\cal A$, and thence its subalgebras ${\cal A}_c$ and $\cal B$, on $\H_1$ as follows: \begin{equation} {\pi}(W({\phi}^{\mu}, {\phi}=0)) = e^{\lbrack{ {\hat{a}_{\mu}(\tilde{\phi}_{\mu})} - {\hat{a}_{\mu}(\tilde{\phi}_{\mu})^*} }\rbrack}, \end{equation} where $\tilde{\phi}_{\mu}=\left( -\overline{\phi}_0, {\phi}_i \right)$, and the symbol ${\pi}$ for a representation has been introduced. The essential point is that the Euclidean commutation relations (\ref{ccr-euclid}) are able to represent the Minkowski commutators (\ref{ccrsk}) because of the special definition of $\tilde{\phi}_{\mu}$. Now, we present a very economical notation for symmetric $n$-particle states by introducing `exponential vectors' \cite{Guichardet}. To this aim, we represent the algebra ${\cal A}_c$ on the dense subset $L_1$ of ${\cal H}_{1}$, which is the span of all exponential vectors \begin{eqnarray} L_1 &=& {\lbrace{\sum_{i=1}^N{\lambda}_{i}e^{{\psi}^{(i)}} \mid {\lambda}_i\in{{\Bbb C}}, {\psi}^{(i)}\in L^2({{\Bbb R}}^3)\otimes {{\Bbb C}}^4, N<\infty}\rbrace}; \nonumber \\ e^{\psi} &:=& 1\oplus\psi \oplus {1\over \sqrt 2}\psi\otimes\psi \oplus {1\over \sqrt{3!}}\psi\otimes\psi\otimes\psi \oplus\ldots , \ll{exp-vect} \end{eqnarray} where the tensor products are understood to be symmetrized. Note that the prefactors $1\over \sqrt{n!}$ of the $n$-particle contributions to $e^{\psi}$ have been chosen differently from those of a Taylor expansion of $e^x$. This allows for a simple form of the scalar product on $L_1$, \begin{equation} (e^{\psi},e^{\varphi}) = e^{(\psi,\varphi)_E}.\ll{e-scalar-prod} \end{equation} A useful remark is now that symmetric $n$-particle states can be obtained from suitably normalized derivatives of exponential vectors, \begin{equation} {\psi}_1{\otimes}_s ... {\otimes}_s {\psi}_n = {1\over \sqrt{n!}} {d\over dr_1} ... {d\over dr_n} e^{\sum_ir_i{\psi}_i} |_{r_i = 0}. \ll{deriv} \end{equation} The representation of $W({\phi}_{\mu},0)$ takes a very simple form on $L_1$. From (\ref{exp-vect}) we have $e^{\hat{a}_{\mu}({\phi}^{\mu})} e^{\psi} = e^{{(\psi,\phi)_E}}e^{\psi}$, $e^{\hat{a}_{\mu}({\phi}^{\mu})^*} e^{\psi} = e^{(\psi+\phi)}$ and hence\footnote{ To see that this defines a representation, we check that $$ {\pi}(W({\phi}_{\mu},0)) {\pi}(W({\varphi}_{\mu},0)) = e^{\lbrack{i{\rm Im}{\lbrack{ (\overline{\phi}_0,\overline{\varphi}_0)_E + ({\phi}_i,{\varphi}_i)_E}\rbrack}}\rbrack} {\pi}(W({\phi}_{\mu}+{\varphi}_{\mu},0)),$$ where $ {\rm Im}{\lbrack{ (\overline{\phi}_0,\overline{\varphi}_0)_E + ({\phi}_i,{\varphi}_i)_E}\rbrack} = B_{sk}({\varphi}_{\mu},0;{\phi}_{\mu},0)$.} \begin{equation} {\pi}(W({\phi}_{\mu},0))e^{\psi} = e^{{-1\over 2}(\phi,\phi)_E+(\psi,\tilde{\phi})_E} e^{(\psi-\tilde{\phi})}. \end{equation} The construction given above is easily generalized to the whole algebra ${\cal A}(M_{sk},B_{sk})$ acting on the dense subspace $ L = L_1 \otimes L_2$ of $ {\cal H} ={\cal H}_1\otimes {\cal H}_2$, where ${\cal H}_2$ is the bosonic Fock space over $L^2({\Bbb R}^3)$. With \begin{equation} L_2 = {\lbrace{ \sum_i^N {\lambda}_i e^{{\psi}^{(i)}} \mid {\psi}^{(i)} \in L^2({{\Bbb R}}^3); {\lambda}_i\in{\Bbb C}, N<\infty }\rbrace} \end{equation} the scalar product of vectors in $L$, reads \begin{equation} (e^{{\psi}_{\mu}}\otimes e^{{\psi}},e^{{\chi}_{\mu}}\otimes e^{{\chi}}) = e^{ ({\psi}_{\mu},{\chi}_{\mu})_E + ({\psi},{\chi})}, \end{equation} and the action of ${\cal A}(M_{sk},B_{sk})$ (denoted by $\pi$ as well, with slight abuse of notation) is \begin{equation} {\pi}(W({\phi}_{\mu},{\phi})) e^{{\psi}_{\mu}}\otimes e^{{\psi}} = e^{{-1\over 2}({\phi}_{\mu},{\phi}_{\mu})_E + ({\psi}_{\mu},\tilde{{\phi}_{\mu}})_E} e^{{-1\over 2}({\phi},{\phi}) + ({\psi},{\phi})} e^{{\psi}_{\mu}- \tilde{{\phi}_{\mu}}} \otimes e^{{\psi}-\phi}. \ll{sk-rep} \end{equation} It should be pointed out that $L$ is only stable under finite linear combinations of the $W$'s (which span a dense subalgebra of $\cal A$), and not under all elements of $\cal A$. Hence, strictly speaking, the induction process is performed relative to the corresponding dense subalgebras of ${\cal A}_c$ and $\cal B$. \subsubsection{Constructing the physical one-particle Hilbert space} With (\ref{sk-rep}), we have specified the bimodule $L$ for ${\cal A}_c$ and $\cal B$, which in this case is a subspace of an `unphysical' Hilbert space ${\cal H}$. Our next step is to construct the corresponding physical Hilbert space, i.e., to carry out the discussion following (\ref{diag2}). In this and the next subsection, we determine the null space ${\cal N}_{sk}$ for the St\"uckelberg-Kibble model, thereby eventually obtaining ${\cal H}_{phys}$. We start from the inner product on elementary vectors in $L$ \begin{equation} (e^{{\psi}_{\mu}}\otimes e^{\psi}, e^{{\chi}_{\mu}}\otimes e^{\chi})_0 = \int_{T_{sk}} [{\cal D}g] ({\pi}(W({\partial}_{\mu}g,-eg)) e^{{\psi}_{\mu}}\otimes e^{\psi}, e^{{\chi}_{\mu}}\otimes e^{\chi}), \ll{pisk} \end{equation} which is a natural generalization of (\ref{0-prod}) (and can, at least heuristically, be derived from an appropriate rigging map defined by a unitary representation of the gauge group on $\cal H$). As in \cite{Landswied}, the heuristic path integral (\ref{pisk}) can be turned into a well-defined integral w.r.t.\ a certain cylindrical measure on $T_{sk}=G$, but here we shall proceed with the formal flat measure ${\cal D}g$, and certify that all manipulations below can be rigorously justified. Using the representation (\ref{sk-rep}) of ${\cal A}(N_{sk},B_{sk})$, we obtain, with $k_0 = \sqrt{e^2 + {\bf k}^2}$, and $d\tilde{k} = {d{\bf k}^3\over ({2\pi})^32k_0}$, \begin{eqnarray} (e^{{\psi}_{\mu}}\otimes e^{\psi}&,& e^{{\chi}_{\mu}}\otimes e^{\chi})_0\nonumber \\ && = e^{\int \tilde{dk} {-1\over k_0^2}{\lbrack{ (k_i{\psi}_i- ie\psi)k_0{\psi}_0 + (k_i\overline{\chi}_i + ie\overline{\chi})k_0\overline{\chi}_0 }\rbrack}}\nonumber \\ &&\times e^{ \int \tilde{dk} {\psi}_i{\left({ {\delta}_{ij} - {k_ik_j\over {\bf k}^2}}\right)} \overline{\chi}_j + {\left({ {e\over k_0}{\psi}_i + i {k_i\over k_0}\psi}\right)} {k_ik_j\over {\bf k}^2} \overline{\left({ {e\over k_0}{\chi}_j + i {k_j\over k_0}\chi}\right)} },\ll{sk-0-prod} \end{eqnarray} where we have used $ {\left({ {\delta}_{ij} - {k_ik_j\over k_0^2}}\right)} = {\left({ {\delta}_{ij} - {k_ik_j\over {\bf k}^2}}\right)} + {e^2 k_ik_j\over k_0^2 {\bf k}^2}$ to write (\ref{sk-0-prod}) in terms of projection operators. To investigate the structure of the null space ${\cal N}_{sk}$, we derive the $(.,.)_0$-inner product for $n$-particle vectors in ${\cal H}$ from (\ref{sk-0-prod}). For one-particle vectors in the (unphysical) space ${\cal H}$, we have \begin{equation} {d\over dr} e^{r{\psi}_{\mu}}\otimes e^{r\psi}|_{r=0} = {\psi}_{\mu} \otimes {\Omega}' + {\Omega}''\otimes\psi, \end{equation} where ${\Omega} = {\Omega}''\otimes {\Omega}'$ denotes the vacuum state in ${\cal H}$. Since such expressions become cumbersome for higher derivatives, for notational convenience we define \begin{equation} {\psi}_*^{(1)}\times ...\times {\psi}_*^{(n)} := {1\over \sqrt{n!}}{d\over dr_1} ... {d\over dr_n} e^{\sum_i r_i {\psi}_{\mu}^{(i)}} \otimes e^{\sum_j r_j {\psi}^{(j)}} {\mid}_{r_i=0}. \end{equation} Then, the $(.,.)_0$-inner product on one-particle vectors in ${\cal H}$ reads \begin{equation} ({\psi}_*,{\chi}_*)_0 = { \int \tilde{dk} {\psi}_i{\left({ {\delta}_{ij} - {k_ik_j\over {\bf k}^2}}\right)} \overline{\chi}_j + {\left({ {e\over k_0}{\psi}_i + i {k_i\over k_0}\psi}\right)} {k_ik_j\over {\bf k}^2} \overline{\left({ {e\over k_0}{\chi}_j + i {k_j\over k_0}\chi}\right)}}. \ll{one-part-prod} \end{equation} Clearly, the two transversal components $P_T{\psi}_* := {\left({ {\delta}_{ij} - {k_ik_j\over {\bf k}^2}}\right)} {\psi}_j$ and a linear combination $P_L{\psi}_*$ of the longitudinal component ${k_ik_j\over{\bf k}^2}{\psi}_j({\bf k})$ with the scalar component ${\psi}({\bf k})$ survive, while the remaining two components lie in ${\cal N}_{sk}$. To be more precise, we introduce for ${\psi}_*$ the {\it Bogoliubov-transformed components} ${\psi}_L$, ${\psi}_N$, \begin{eqnarray} {\psi}_{L,i}({\bf k}) &:=& \cos\theta {k_ik_j{\psi}_j({\bf k})\over {\bf k}^2} + i\sin\theta {k_i{\psi}({\bf k})\over |{\bf k}|} ,\nonumber \\ {\psi}_{N,i}({\bf k}) &:=& -\sin\theta {k_ik_j{\psi}_j({\bf k})\over {\bf k}^2} + i\cos\theta {k_i{\psi}({\bf k})\over |{\bf k}|}, \end{eqnarray} where $\cos\theta = {e\over k_0}$, $\sin\theta = {|{\bf k}|\over k_0}$. With ${\psi}_L$, ${\psi}_T$ and ${\psi}_N$, the five-component vector ${\psi}_*^{(i)}$ can be specified as \begin{equation} {\psi}_*({\bf k}) := \left({P_T{\psi}_{\mu}({\bf k}), {\psi}_L({\bf k}), {\psi}_N({\bf k}), {\psi}_0({\bf k})}\right), \end{equation} and the projection operator $P_p$ onto the `physical' one-particle components is given by \begin{equation} (P_p{\psi}_*)({\bf k}) = \left({P_T{\psi}_{\mu}({\bf k}), {\psi}_L({\bf k}),0,0}\right). \ll{one-part-phys} \end{equation} This is exactly what one expects: the five `unphysical' degrees of freedom have combined into three physical ones in such a way that the longitudinal component in ${\cal H}$ has mixed with the scalar component. \subsubsection{The physical Hilbert space ${\cal H}_{phys}$} To extend (\ref{one-part-phys}) to $n$-particle states, we rewrite (\ref{sk-0-prod}), using $$\exp(\sum_ir_i{\psi}_*^{(i)}) := \exp(\sum_ir_i{\psi}_{\mu}^{(i)}) \otimes \exp(\sum_ir_i{\psi}^{(i)}),$$ \begin{equation} (e^{{\psi}_*},e^{{\chi}_*})_0 = (e^{{\psi}_*},\Omega)_0 (\Omega, e^{{\chi}_*})_0 (e^{P_p{\psi}_*},e^{P_p{\chi}_*}). \ll{nullsk} \end{equation} Here we have used the remark following (\ref{one-part-prod}), which implies that $$ (\exp(P_p{\psi}_*),\exp(P_p{\chi}_*))_0 = (\exp(P_p{\psi}_*),\exp(P_p{\chi}_*)).$$ {}From (\ref{nullsk}) we obtain \begin{eqnarray} {\psi}_*^{(1)}\times ... \times{\psi}_*^{(n)} &=& {d\over dr_1}...{d\over dr_n} (e^{\sum_ir_i{\psi}_*^{(i)}},\Omega)_0 e^{\sum_ir_i{\psi}_*^{(i)}} |_{r_i=0}\nonumber \\ &=& \sum_{q=0}^n \sum_{(p_i)_1^q\in {\cal P}_{q,n}} {\lambda}_{(p_i)_1^q} (P_p{\psi}_*^{(p_1)}){\times} ... {\times}(P_p{\psi}_*^{(p_q)}) + \vec{n},\ll{n-part-unphys} \end{eqnarray} where ${\cal P}_{q,n}$ contains all sets of $q$ indices ${\lbrace{(p_i)_1^q}\rbrace}$ out of ${\lbrace{1,...,n}\rbrace}$, such that ${\lbrace{(p_i)_1^q}\rbrace}$ $\cup$ ${\lbrace{(\hat{p_i})_1^{n-q}}\rbrace}$ $={\lbrace{1,...,n}\rbrace}$. Here, \begin{equation} {\lambda}_{(p_i)_1^q} = \sqrt{(q)! (n-q)!\over n!} ({\psi}_*^{(\hat{p}_1)}\times ... \times {\psi}_*^{(\hat{p}_{n-q})}{\mid}_{p_q(I_{n,q})},\Omega)_0 \end{equation} are $c$-number coefficients and $\vec{n}$ denotes an element in ${\cal N}_{sk}$. Vectors of the type (\ref{one-part-phys}) generate a Hilbert space of physical one-particle states. The bosonic Fock space over this one-particle space is evidently ${\cal F}_{phys}:={\cal S}(L^2({{\Bbb R}}^3)\otimes {{\Bbb C}}^3)$, the symmetric Hilbert space over $(L^2({{\Bbb R}}^3)\otimes {{\Bbb C}}^3)$. It should be clear from equation (\ref{n-part-unphys}) that the induced space ${\cal H}_{phys}$ from the Rieffel induction procedure is naturally isomorphic to this physical Fock space.\footnote{ Of course, all Hilbert spaces of the same dimension are unitarily equivalent, but to impose such equivalence one generally has to pick a basis. We use the term `naturally isomorphic' to indicate that a unitary equivalence exists which doesn't require the choice of a basis. {}From the point of view of representation theory, this equivalence intertwines the actions of appropriate operator algebras, cf.\ the next subsection.} To prove this, we define a map $V:L\rightarrow {\cal F}_{phys}$ by linear extension of $V\exp(\psi_*)=(\exp(\psi_*),\Omega)_0\exp(P_p\psi_*)$. It follows from an argument similar to the one in section 3.3 of \cite{Landswied} that this map is well-defined (which is a nontrivial property, as the basis $\{\exp(\psi_*)\}$ is overcomplete). Eq.\ (\ref{nullsk}), and the fact that the inner product in ${\cal F}_{phys}$ is just the one in $\cal H$, restricted to the physcial states, then implies the crucial property \begin{equation} (V\Psi,V\Phi)=(\Psi,\Phi)_0 \ll{crucial} \end{equation} for all $\Psi,\Phi\in L$, where the inner product on the l.h.s.\ is evidently the one in ${\cal F}_{phys}$. Hence the null space ${\cal N}_{sk}$ of $(.,.)_0$ is precisely the kernel of $V$, and the quotient map $\tilde{V}: L/{\cal N}_{sk}\rightarrow {\cal F}_{phys}$ can be extended to a unitary map (denoted by the same symbol) $\tilde{V}: {\cal H}_{phys}\rightarrow {\cal F}_{phys}$. \subsubsection{$n$-point correlation functions and gauge-invariance} Having specified the physical Hilbert space ${\cal H}_{phys}$, the next step is to determine the action of ${\pi}_{phys}({\cal A}_c)$. To this end, we consider the generating functional ${\omega}_{vac}$ for vacuum expectation values, \begin{eqnarray} {\omega}_{vac}({\phi}^{\mu},\phi) &:=& ({\pi}(W({\phi}_{\mu},\phi))\Omega, \Omega)_0 \nonumber \\ &=& e^{{1\over 2}({\phi}_{\mu},{\phi}_{\mu})_M} e^{-{1\over 2}({\phi},{\phi})} e^{-{1\over k_0^2}(k_0\overline{\phi}_0(k_{\mu}{\phi}_{\mu} + ie\phi))}, \ll{vac-exp} \end{eqnarray} where $\Omega \in {\cal H}$ is the (unphysical) `vacuum' state. By construction, only ${\cal A}_c={\cal A}(N_{sk},B_{sk})$ acts on ${\cal H}$ (cf.\ (\ref{nsk})), and for $({\phi}_{\mu},{\phi}) \in N_{sk}$, $k_0{\phi}_0 = k_i{\phi}_i - ie{\phi}$, we obtain \begin{eqnarray} {\omega}_{vac}({\phi}^{\mu},\phi) &=& e^{-{1\over 2}({\phi}_{\mu},P_T{\phi}_{\mu})_E} e^{-{1\over 2}({e\over k_0}{\phi}_i + i{k_i\over k_0}\phi) {k_ik_j\over {\bf k}^2} ({e\over k_0}\overline{\phi}_j - i{k_i\over k_0}\overline{\phi})}\nonumber \\ &=:& ({\pi}_{phys}(\tilde{W}(P_p{\phi}_*)){\Omega}_{phys}, {\Omega}_{phys})_{phys} \nonumber \\ &=& e^{-\mbox{\footnotesize $\frac{1}{2}$}\int d\tilde{k} [\overline{P_T\phi({\bf k})}P_T\phi({\bf k}) + \overline{\phi}_{L,i}({\bf k}){\phi}_{L,i}({\bf k})]}. \end{eqnarray} Here, ${\Omega}_{phys} \in {\cal H}_{phys}$ is the physical vacuum state; it is just the projection of $\Omega\in L$ onto $L/{\cal N}_{sk}\, \subset {\cal H}_{phys}$. We observe that for $({\phi}_{\mu},\phi)\in T_{sk}$, ${\pi}(W({\phi}_{\mu},\phi))$ equals the unit operator, cf.\ (\ref{tsk}). This implies that the gauge group is represented trivially on ${\cal H}_{phys}$. Moreover, one infers that ${\cal A}_{obs}:= {\pi}({\cal A}_c)\simeq {\cal A}(N_{sk}/T_{sk},B_{c,sk})$, since the image of a representation of a $C^*$-algebra is isomorphic to the algebra quotiented by the kernel of the representation. Now $N_{sk}/T_{sk}\simeq P_p N_{sk}$ as vector spaces (but not as carrier spaces of actions of the Poincar\'{e} group!), so that, equally well, ${\cal A}_{obs}\simeq {\cal A}(P_p N_{sk},B_{sk})$.\footnote{ However, the isomorphism between ${\cal A}(P_p N_{sk},B_{sk})$ and ${\cal A}(N_{sk}/T_{sk})$ does not preserve the (automorphic) action of the Poincar\'{e} group, which, indeed, acts on the latter but not on the former, cf.\ \cite{CGH}.} Then, it is clear from section 3.1 that ${\cal A}_{obs}$ is precisely the Weyl algebra over de Marsden-Weinstein reduced space (i.e., the physical phase space) of the St\"uckelberg-Kibble model. Hence it describes three gauge-invariant, massive field components. Thus $\tilde{W}(P_p{\phi}_*))$ can be viewed as a Weyl operator in ${\cal A}(P_pN_{sk},B_{sk})$. In particular, the representation of ${\cal A}(P_pN_{sk},B_{sk})$ on exponential vectors $e^{\psi} \in {\cal H}_{phys} = {\cal S}(L^2({{\Bbb R}}^3)\otimes {{\Bbb C}}^3)$ is given by \begin{eqnarray} {\pi}_{phys}(\tilde{W}(P_p{\phi}_*))e^{\psi} = e^{-\mbox{\footnotesize $\frac{1}{2}$}(P_p{\phi}_*,P_p{\phi}_*)_p + (\psi,P_p{\phi}_*)_p} e^{(\psi - P_p{\phi}_*)} \nonumber \\ (\psi,P_p{\phi}_*)_p = \int d\tilde{k} [\overline{P_T\phi({\bf k})}P_T\phi({\bf k}) + \overline{\phi}_{L,i}({\bf k}){\phi}_{L,i}({\bf k})]. \end{eqnarray} {}From ${\omega}_{vac}({\phi}_{\mu},{\phi})$, $n$-point correlation functions can be obtained as multiple derivatives of $\tilde{W}(P_p{\phi}_*) := e^{i\tilde{A}(f)}$, where $P_p{\phi}_* = \triangle * f \in L^2({{\Bbb R}}^3)\otimes {{\Bbb C}}^3$. \begin{eqnarray} &i^n& ({\pi}_{phys}(\tilde{A}(f_1)...\tilde{A}(f_n){\Omega}_{phys}, {\Omega}_{phys})_{phys} = {d\over dr_1}...{d\over dr_n} {\omega}_{vac}(\sum_ir_i{\phi}_{\mu}^{(i)}, \sum_ir_i{\phi}^{(i)})|_{r_i=0}\nonumber \\ &=& \sum_{(p_i,q_i)_i^{n\over 2} \in {\cal S}_n} \prod_{i=1}^{n\over 2} ({\pi}_{phys}(\tilde{A}(f_{p_i})\tilde{A}(f_{q_i}){\Omega}_{phys}, {\Omega}_{phys})_{phys}(-1)^{n\over 2} \ll{n-point} \end{eqnarray} for $n$ even and zero otherwise. Here, ${\cal S}_n$ denotes the set of all symmetric partitions of ${\lbrace{1, ...,n}\rbrace}$ into a set of unordered pairs $(p_i,q_i)$. We conclude from (\ref{n-point}) that the $n$-point correlation functions can be decomposed into products of $2$-point correlation functions, i.e., Wick's theorem is satisfied. The reader should note, however, that this form of Wick's theorem is satisfied for elements in ${\cal A}(N_{sk},B_{sk})$ only. The crucial point is that in general, the $(.,.)_0$-inner product preserves the adjoint for test functions in $N_{sk}$ only. This can be seen by comparing, e.g., ${d\over dr_1}{d\over dr_2} ({\pi}(W(\sum_ir_i{\phi}_{\mu}^{(i)},$ $\sum_ir_i{\phi}^{(i)}))\Omega,\Omega)_0 |_{r_i=0}$ with ${d\over dr_1}{d\over dr_2} ({\pi}(W({\phi}_{\mu}^{(1)},$ ${\phi}^{(1)}))\Omega, {\pi}(W({\phi}_{\mu}^{(2)}, {\phi}^{(2)}))\Omega)_0 |_{r_i=0}$, cf. (\ref{vac-exp}). There is an interesting parallel between this restriction of the Rieffel induced expectation values to ${\cal A}(N_{sk},B_{sk})$ and the general set-up of the Gupta-Bleuler indefinite metric formalism as presented in \cite{Strocchi}. In the latter, one starts from an unphysical Hilbert space ${\cal H}_{GB}$ from which the physical one is obtained as a quotient ${\cal H}'/{\cal H}''$. Without reviewing this construction, we note that ${\cal H}$ has to be restricted to a suitable subspace ${\cal H}' \subset {\cal H}_{GB}$ before quotiening by a null space ${\cal H}'$. Obviously, in our setting, a similar restriction is needed on the level of the algebra, ${\cal A}(N_{sk},B_{sk})\subset {\cal A}(M_{sk},B_{sk})$. This restriction emerges in a systematic way, for as we pointed out before, the subalgebra in question is the commutant of the algebra generated by the constraints (i.e., by the gauge group). This observation is closely related to the result of Narnhofer and Thirring \cite{Narnhofer} that covariant formulations without indefinite inner metric are possible as long as the representation on the physical Hilbert space is restricted to a certain subalgebra of weak observables. In the example of Narnhofer and Thirring, non-regular states have to be introduced. This can be avoided in the Rieffel induction setting, cf. \cite{Landswied,UAW} for further details. \subsubsection{Positivity of the Hamiltonian and action of the Poincar\'e group} On the algebra of weak observables of the St\"uckelberg-Kibble model ${\cal A}(N_{sk},B_{sk})$, the time evolution is given as an automorphism group ${\tau}_t$, \begin{equation} {\tau}_t[W({\phi}_{\mu},\phi)] = W(e^{it\sqrt{D+e^2}}{\phi}_{\mu}, e^{it\sqrt{D+e^2}}\phi ), \end{equation} where $(D{\phi})_{\mu} = (-{\triangle}{\phi}_0, -{\triangle}{\phi}_1, -{\triangle}{\phi}_2, -{\triangle}{\phi}_3)$. We want to construct the Hamiltonian $H$, corresponding to ${\tau}_t$ on ${\cal H}$. $H$ is a representation-dependent operator, implementing the time evolution ${\tau}_t$ in the representation ${\pi}$ by \begin{equation} e^{itH}{\pi}(W({\phi}_{\mu},\phi)) e^{-itH} = {\pi}({\tau}_t[W({\phi}_{\mu},\phi)]). \end{equation} Comparing this with the explicit form of the representation in terms of annihilation and creation operators $\hat{a}_{\mu}^*$, $\hat{a}_{\mu}$ for the vector field and $\hat{b}^*$, $\hat{b}$ for the scalar field, we obtain \begin{equation} H = - \int d\tilde{k} \sqrt{{\bf k}^2+e^2} \hat{a}_{\mu}^*({\bf k}) g^{\mu\nu} \hat{a}_{\nu}^({\bf k}) + \int d\tilde{k} \sqrt{{\bf k}^2+e^2} \hat{b}^*({\bf k}) \hat{b}({\bf k}). \end{equation} Regarded as an operator on $\cal H$ (with its Hilbert space inner product), this Hamiltonian clearly has the entire real axis as its spectrum. However, it is easy to see that \begin{equation} ({\Psi}, H {\Psi})_0 \geq 0 \end{equation} for all $\Psi \in {\cal H}$. The point is that arbitrary (normalized) components of the physical one-particle state space, ${\left({ {\delta}_{ij} - {k_ik_j\over {\bf k}^2}}\right)}{\psi}_j$ and ${k_i\over {\bf k}}{\psi}_i \cos\theta + i\psi\sin\theta$ pick up (the same) positive energy contributions. For multi-particle states, this holds true due to their decomposition into such components. The elements of ${\cal H}$ carrying the negative energy spectrum have ended up in the null space. Hence the induced Hamiltonian $H_{phys}$ on ${\cal H}_{phys}$ is positive. Finally, we note that ${\cal H}_{phys}$ carries a massive representation of the Poincar\'e group ${\cal P}$. Indeed, ${\omega}_{vac}$ is Poincar\'e invariant on $N_{sk}$ and hence \cite{Haag,BW} there exists a Poincar\'e invariant vacuum state ${\Omega}_{phys}\in {\cal H}_{phys}$ and a representation $U_p$ of the Poincar\'e group, such that \begin{equation} U_p(\Lambda,a){\pi}_{phys}(W({\phi}_{\mu},\phi)){\Omega}_{phys} = {\pi}_{phys}(W({\gamma}_{\Lambda,a}({\phi}_{\mu},\phi))){\Omega}_{phys} \end{equation} for all $({\phi}_{\mu},\phi)\in N_{sk}$. It is easily shown that $H_{phys}$ is the generator of the time-translation part of the representation thus defined. Since the spectrum of the Hamiltonian $H_{phys}$ shows a mass gap, we are dealing with a massive representation ($m^2 = e^2$) of the Poincar\'e group, i.e., the three components of the vector $P_p{\psi}_*^{(i)}$ transform as a massive one-particle state under the action of the little group $SO(3)$ \cite{Barut}. We conclude that ${\cal H}_{phys}$ has the main properties required by a physical Hilbert space: it transforms trivially under the gauge group, satisfies the positive spectrum condition and carries a unitary representation of the Poincar\'e group. \section{Conclusion} The quantization proposal employed in this paper provides a detailed scheme for imposing constraints on gauge quantum field theories. As explained in Chapter 2, the main tool of this proposal is the Rieffel induction procedure, which provides a systematic scheme for the construction of representations of $C^*$-algebras. It may be viewed as the quantum counterpart of the symplectic reduction technique; as we have shown, this is particulary obvious for Weyl $C^*$-algebras. This leads to a new quantization method for gauge field theories. In the present work, we have applied this method to the St\"uckelberg-Kibble model. To this end, we have defined a field algebra ${\cal A}$ corresponding to the field content of the Lagrangian, and an algebra of constraints ${\cal B}$ corresponding to the gauge group acting on ${\cal A}$. Also, we have specified a representation ${\pi}$ of subalgebras of ${\cal A}$ on a (unphysical) Hilbert space ${\cal H}$. From these input data, we have constructed a representation of the physical, gauge-invariant fields on a new Hilbert space ${\cal H}_{phys}$. The construction of ${\cal H}_{phys}$ shows some parallels to the Gupta-Bleuler indefinite metric formalism. In both settings, a degenerate inner product is defined on a (unphysical) Hilbert space ${\cal H}$, and ${\cal H}_{phys}$ is constructed by quotiening ${\cal H}$ by a null space with respect to this degenerate inner product. Yet, there are important differences. In contrast to the indefinite metric inner product ${\langle{.,.}\rangle}$, defined on ${\cal H}_{GB}$ in the Gupta-Bleuler formalism, the $(.,.)_0$-inner product is positive semidefinite. More importantly, it is a conceptual advantage of our quantization method that $(.,.)_0$ is derived from first principles (namely from the requirement to impose quantum constraints by a quantized version of the classical phase space reduction method), whereas the Gupta-Bleuler formalism takes ${\langle{.,.}\rangle}$ as starting point without further justification. A similar comment applies to the BRST technique: although a classical analogue of this procedure exists, the quantum BRST procedure is {\em not} in any satisfactory sense the quantization of the classical scheme. Another remarkable difference between both formalisms is that the Gupta-Bleuler formalism restricts the unphysical Hilbert space before forming the quotient while the proposal of \cite{Landsman} restricts itself to a representation of the subalgebra ${\cal A}_c$ of weak observables on ${\cal H}$, before quotiening by the appropriate null space. As a consequence, the $(.,.)_0$-inner product preserves the adjoint for elements in ${\cal A}_c$ only. It remains to be seen how far this feature alters applications of usual perturbative techniques in more complicated models. Most of our effort in Chapter 3 has gone into characterizing the particular features of the physical state space ${\cal H}_{phys}$. By construction, ${\cal H}_{phys}$ carries a trivial representation of the gauge group. Also, the states are physical in the sense that they obey a positive spectrum condition and that they carry a massive representation of the Poincar\'e group. Since the St\"uckelberg-Kibble model has been widely used in investigations of the Higgs mechanism, we emphasize again the result obtained for the one-particle subspace in ${\cal H}_{phys}$. The point is that in our proposal, the particular construction method of ${\cal H}_{phys}$ allows one to trace back how the (unphysical) components of ${\cal H}$ end up in the physical Hilbert space. In the present case, we have shown that the longitudinal physical one-particle component arises from a particular Bogoliubov-transformation of the unphysical longitudinal and the scalar component. As expected from general considerations, two of the five components in ${\cal H}$ have ended up in the one-particle null space. We conclude our discussion of the Rieffel induction procedure by pointing out that our presentation has focused on a particular way of applying Rieffel induction to gauge quantum field theories. Conceptually, the scheme is much wider. It remains to be seen how far other choices for the inner product $(.,.)_0$ and the unphysical Hilbert space ${\cal H}$ allow for other realisations of the physical observables of gauge field theories.
\section{INTRODUCTION} \bigskip In previous papers [1--4] the Hamilton--Jacobi formulation of singular systems has been studied. This formulation leads us to the following total differential equations: \setcounter{equation}{0} \renewcommand{\theequation}{1.\arabic{equation}} \begin{equation} dq_a={\partial H'_\alpha\over\partial p_a}\ dx_\alpha ,\quad dp_a=-{\partial H'_\alpha\over\partial q_a}\ dx_\alpha,\quad dp_\alpha =-{\partial H'_\beta\over\partial x_\alpha}\ dx_\beta \end{equation} $$ \alpha ,\beta =0,1,\dots , r;\qquad a=1,2,\dots , n-r\ $$ with constraints \begin{equation} H'_\alpha =H_\alpha (x_\beta , q_a,p_a)+p_\alpha \end{equation} (Note that we are adopting summation convention in this work.) Solutions of Eqs.(1.1) give the field, $q_a$, in terms of independent coordinates \begin{equation} q_a\equiv q_a(t,x_\mu ),\qquad \mu =1,2,3,\dots ,r \end{equation} where $x_0=t$. The link between the Hamilton--Jacobi approach and the Dirac approach is studied in Ref.[5]. In Ref.[6] the singular Lagrangians are treated as continuous systems. The Euler--Lagrange equations of constrained systems are proposed in the form \begin{equation} {\partial\over\partial x_\alpha}\ \left[ {\partial L'\over \partial (\partial_\alpha q_a)}\right] -{\partial L'\over\partial q_a}=0;\quad \partial_\alpha q_a={\partial q_a\over\partial x_\alpha}\ . \end{equation} with constraints \begin{eqnarray} dG_0&=&-{\partial L'\over\partial t}\ dt\\ dG_\mu &=& -{\partial L'\over\partial x_\mu}\ dt \end{eqnarray} where \begin{equation} L'(x_\alpha ,\partial_\alpha q_a,\dot x_\mu ,q_a)\equiv L(q_a,x_\alpha ,\dot q_a=(\partial_\alpha q_a)\dot x_\alpha ) \end{equation} $$ \dot x_\mu ={dx_\mu\over dt};\qquad \dot x_0=1 $$ \begin{equation} G_\alpha =H_\alpha \left( q_a,x_\beta ,p_a={\partial L\over\partial \dot q_a}\right) \end{equation} The variation of constraints (1.5,6) should be taken into consideration in order to have a consistent theory. An instructive work is the canonical formalism for degenerate Lagrangians [7--9]; the starting point of this formalism is to consider Lagrangians with ranks of the Hessian matrix less than $n$. Shanmugadhasan has called these systems as degenerate. For such systems some of the Euler--Lagrange equations do not contain acceleration. Following Refs.[7--9] these equations are considered as constraints. In other words, if the rank of Hessian matrix is $(n-r)$, with $r<n$, then the Euler--Lagrange equations can be expressed in the form \begin{eqnarray} {d\over dt}\ \left({\partial L\over\partial\dot q_a}\right) -{\partial L\over\partial q_a} &=& 0\\ {d\over dt}\ \left({\partial L\over\partial\dot x_\mu}\right) -{\partial L\over\partial x_\mu} &=& 0 \end{eqnarray} With the aid of Eq.(1.9), Eq.(1.10) can be identically satisfied, i.e. $0=0$, or they lead to equations free from acceleration. These equations are diveded into two types: type--$A$ which contains coordinates only and type--$B$ which contains coordinates and velocities [9]. The total time derivative of the above two types of constraints should be considered in order to have a consistent theory. In this paper we would like to study the link between the treatment of singular Lagrangians as field systems [6] and the well--known Lagrangian formalism. In Section 2 the relation between the two approaches is discussed, and in Section 3 two examples of singular Lagrangians are constructed and solved using the two approaches. In Section 4 the treatment of a singular Lagrangian with Hessian matrix of zero rank is discussed. \section{THE RELATION BETWEEN THE TWO APPROACHES} One should notice that Eqs.(1.4) are equivalent to Eqs.(1.9). In other words the expressions \setcounter{equation}{0} \renewcommand{\theequation}{2.\arabic{equation}} \begin{equation} (\partial_\alpha\ q_a)\ \dot x_\alpha \end{equation} and \begin{equation} \partial_\beta (\partial_\alpha\ q_a\ \dot x_\alpha )\ \dot x_\beta \end{equation} In Eqs.(1.4) can be be replaced by $\dot q_a$ and $\ddot q_a$ respectively in order to obtain Eqs.(1.9). Following Refs.[1--6], we have \begin{equation} G_0\equiv H_0\ , \end{equation} and \begin{equation} G_\mu\equiv H_\mu =-p_\mu \end{equation} Thus, Eqs.(1.6) lead to \begin{equation} {dp_\mu\over dt}={\partial L\over\partial x_\mu} \end{equation} Making use of the definition of momenta, Eqs.(2.5) lead to Eqs.(1.10). Hence Eqs.(1.5,6) are equivalent to Eqs.(1.9,10). \section{EXAMPLES} The procedure described in Section 2 will be demonstrated by the following examples.\\ \\ A.\quad Let us consider a Lagrangian of the form \setcounter{equation}{0} \renewcommand{\theequation}{3.\arabic{equation}} \begin{equation} L={1\over 2}\ \dot q^2_1+\dot q_1\dot q_2+{1\over 2}\ \dot q^2_2+4q_1\dot q_2+(2q^2_1+4q_1q_2) \end{equation} The Euler--Lagrange equations then read as \begin{eqnarray} &&\ddot q_1+\ddot q_2-4\dot q_2-4(q_1+q_2)=0\\ &&\ddot q_1+\ddot q_2+4\dot q_1-4q_1=0 \end{eqnarray} Substituting Eq.(3.2) in Eq.(3.3), gives a $B$--type constraint \begin{equation} B_1=\dot q_2+\dot q_1+q_2=0 \end{equation} For consistent theory, the time derivative of Eq.(3.4) should be equal to zero. This leads to the new $B$--type constraint \begin{equation} B_2=5\dot q_2+4(q_2+q_1)=0 \end{equation} Taking the time derivative for the new constraints we get a second order differential equation for $q_2$ \begin{equation} 5\ddot q_2-4q_2=0 \end{equation} which has the following solution $$ q_2=A\ e^{{2\over\sqrt 5}\ t} +B\ e^{-{2\over\sqrt 5}\ t} $$ Now, let us look at this Lagrangian as a field system. Since the rank of the Hessian matrix is one, the above Lagrangian can be treated as a field system in the form \begin{equation} q_1=q_1(t,q_2);\qquad x_2=q_2 \end{equation} Thus, the expression \begin{equation} \dot q_1={\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2 \end{equation} can be replaced in Eq.(3.1) to obtain the following modified Lagrangian $L'$: \begin{eqnarray} L' &=& {1\over 2}\ \left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2\right]^2+\left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2\right]\ \dot q_2\nonumber\\ &&+{1\over 2}\ \dot q^2_2+4q_1\dot q_2+(2q^2_1+4q_1q_2) \end{eqnarray} Making use of Eqs.(1.4), we have \begin{equation} {\partial^2q_1\over\partial t^2}+2{\partial^2q_1\over\partial t\ \partial q_2}\ \dot q_2+{\partial q_1\over\partial q_2}\ \ddot q_2+{\partial^2q_1\over\partial q^2_2}\ \dot q^2_2+\ddot q_2-4\dot q_2-4(q_1+q_2)=0 \end{equation} Note that we have made the substitution $\alpha =0,2$ and $a=1$, in order to get the above equation. Making use of Eq.(3.8) and the fact that \begin{equation} \ddot q_1={\partial^2q_1\over\partial t^2}+2{\partial^2q_1\over\partial t\ \partial q_2}\ \dot q_2+{\partial q_1\over\partial q_2}\ \ddot q_2+{\partial^2q_1\over\partial q^2_2}\ \dot q^2_2 \end{equation} Eq.(3.10) will be the same as Eq.(3.2). According to Refs.[1--4] the quantity $H_2$ can be calculated as \begin{equation} H_2=-(\dot q_1+\dot q_2+4q_1) \end{equation} Hence, \begin{equation} G_2=-\left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2\right] -\dot q_2-4q_1 \end{equation} and taking the total differential of Eq.(3.13) one gets, \begin{eqnarray} dG_2 &=&-\Biggl\{\left[{\partial^2q_1\over\partial t^2}+2{\partial^2q_1\over\partial q_2\ \partial t}\ \dot q_2+{\partial^2 q_1\over\partial q^2_2}\ \dot q_2+{\partial q_1\over\partial q_2}\ \ddot q_2\right] +\Biggr.\nonumber \\ &&\Biggl. +4\left[{\partial q_1\over\partial t} +{\partial q_1\over\partial q_2}\ \dot q_2\right] +\ddot q_2\Biggr\}\ dt \end{eqnarray} Replacing the expression in the first parenthesis from Eq.(3.10) one gets \begin{equation} dG_2=-\left\{ 4\dot q_2+4[q_1+q_2]+4\left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2\right]\right\}\ dt \end{equation} Making use of Eq.(1.6), one finds \begin{equation} dG_2=-4q_1\ dt \end{equation} Equating Eq.(3.16) with Eq.(3.15), we have the following constraint \begin{equation} F_1=\dot q_2+q_2+{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2=0 \end{equation} Using the expression (3.8), one observes that this constraint is equivalent to the $B$--type constraint (3.4). For a valid theory, the variation of $F_1$ should be zero; thus one gets \begin{equation} dF_1=F_2\ dt=0 \end{equation} where \begin{equation} F_2=5\dot q_2+4q_1+4q_2=0 \end{equation} This is $B_2$ constraint defined in Eq.(3.5). Again taking the total differential of the new constraint $F_2$, we have \begin{equation} dF_2=[5\ddot q_2-4q_2]\ dt=0 \end{equation} Thus \begin{equation} 5\ddot q_2-4q_2=0 \end{equation} This is a second order differential equation for $q_2$ and is the same as Eq.(3.6). In addition, the function $G_0$ can be evaluated and \begin{equation} G_0={1\over 2}\ \left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2\right]^2+{1\over 2}\ \dot q^2_2+\left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2\right]\ \dot q_2-2q^2_1-4q_1q_2 \end{equation} where \begin{equation} dG_0=4\dot q_2\ F_1\ dt \end{equation} and this does not lead to any further constraints.\\ \\ B.\quad Consider the Lagrangian of the form \begin{equation} L={1\over 2}\ (\dot q^2_1+\dot q^2_2)+\dot q_1\dot q_2+{1\over 2}\ (q^2_1+q^2_2) \end{equation} Then, the Euler--Lagrange equations are given as \begin{eqnarray} \ddot q_1+\ddot q_2-q_1 & = & 0\\ \ddot q_1+\ddot q_2-q_2 & = & 0 \end{eqnarray} Expressing Eq.(3.25) as \begin{equation} q_1=\ddot q_1+\ddot q_2 \end{equation} and substituting in Eq.(3.26), one gets an $A$--type constraint \begin{equation} A_1=q_1-q_2=0 \end{equation} There are no further constraints. Thus Eq.(3.26) takes the form \begin{equation} 2\ddot q_2-q_2=0 \end{equation} As in the previous example, this system can be treated as field system, and the modified Lagrangian $L'$ reads as \begin{equation} L'={1\over 2}\left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2\right]^2+{1\over 2}\ \dot q^2_2+\left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2\right]\ \dot q_2+{1\over 2}\ (q^2_1+q^2_2) \end{equation} The Euler--Lagrange equation for this field system is obtained as \begin{equation} {\partial^2q_1\over\partial t^2}+2{\partial^2q_1\over\partial t\ \partial q_2}\ \dot q_2+{\partial q_1\over\partial q_2}\ \ddot q_2+{\partial^2q_1\over\partial q^2_2}\ \dot q^2_2+\ddot q_2-q_1=0 \end{equation} Again replacing $\ddot q_1$ by the expression (3.11), Eq.(3.31) will be the same as Eq.(3.25). Besides, the function $G_2$ can be calculated as \begin{equation} G_2=-\left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2\right] -\dot q_2 \end{equation} and the total differential of $G_2$ can be written as \begin{equation} dG_2=-\left[{\partial^2q_1\over\partial t^2}+2{\partial^2q_1\over\partial t\ \partial q_2}\ \dot q_2+{\partial^2q_1\over\partial\dot q^2_2}\ \dot q^2_2+{\partial q_1\over\partial q_2}\ \ddot q_2+\ddot q_2\right]\ dt \end{equation} Using Eq.(3.31) and Eq.(1.6), we have \begin{equation} dG_2=-q_1\ dt=-q_2\ dt \end{equation} which leads to the following constraint \begin{equation} F_1=q_1-q_2=0 \end{equation} This is an $A$--type constraint of the form (3.28). Taking the total differential of $F_1$, we have \begin{equation} dF_1=\left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2-\dot q_2\right]\ dt=0 \end{equation} and this leads to a new constraint \begin{equation} F_2={\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2-\dot q_2=0 \end{equation} which is equivalent to the total time derivative of the constraint (3.28). Again calculating the total differential of $F_2$, one gets \begin{equation} dF_2=\left[{\partial^2q_1\over\partial t^2}+2{\partial^2q_1\over\partial t\ \partial q_2}\ \dot q_2+{\partial^2q_1\over\partial q^2_2}\ \dot q^2_2+{\partial q_1\over\partial q_2}\ \ddot q_2-\ddot q_2\right]\ dt=0 \end{equation} and making use of Eqs.(3.31) and (3.35), we get \begin{equation} 2\ddot q_2-q_2=0 \end{equation} which is the same as Eq.(3.29). Besides the function $G_0$ is calculated as \begin{equation} G_0={1\over 2}\left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2\right]^2+{1\over 2}\ \dot q^2_2+\left[{\partial q_1\over\partial t}+{\partial q_1\over\partial q_2}\ \dot q_2\right]\ \dot q_2-{1\over 2}\ [q^2_1+q^2_2] \end{equation} Thus, the total differential of $G_0$ is obtained as \begin{equation} dG_0=\dot q_2\ F_1\ dt=0 \end{equation} and with the aid of Eq.(3.35), it is identically satisfied. \section{A SINGULAR LAGRANGIAN WITH ZERO RANK HESSIAN MATRIX} According to the treatment of singular Lagrangians as field systems: if the Hessian matrix has rank equal to zero, the Lagrangian cannot be treated as field system. Whereas, the equation of motion which does not contain acceleration can be obtained using the constraints (1.6). As an example let us consider the following Lagrangian which was given in Ref.[10]: \setcounter{equation}{0} \renewcommand{\theequation}{4.\arabic{equation}} \begin{equation} L=(q_2+q_3)\ \dot q_1+q_4\ \dot q_3+{1\over 2}\ (q^2_4-2q_2q_3-q^2_3) \end{equation} The momenta are obtained as \begin{equation} p_1=q_2+q_3,\quad p_2=0,\quad p_3=q_4,\quad p_4=0 \end{equation} Thus, \begin{eqnarray} G_1 &=& -(q_2+q_3)\\ G_2 &=& 0\\ G_3 &=& -q_4\\ G_4 &=& 0 \end{eqnarray} Making use of Eqs.(1.6), one gets \begin{eqnarray} dG_1 &=& -(\dot q_2+\dot q_3)\ dt=0\\ dG_2 &=& 0=-(\dot q_1-q_3)\ dt\\ dG_3 &=& -\dot q_4\ dt =-(\dot q_1-q_2-q_3)\ dt\\ dG_4 &=& 0=-(\dot q_3+q_4)\ dt \end{eqnarray} These equations lead to the following equations of motion \begin{eqnarray} \dot q_2+\dot q_3 &=& 0\\ \dot q_1-q_3 &=& 0\\ \dot q_4-\dot q_1+q_2+q_3 &=& 0\\ \dot q_3+q_4 &=& 0 \end{eqnarray} and these are the Euler--Lagrange equations which are free from acceleration, and are of $B$--type constraints. \section{CONCLUSIONS} As it was mentioned in the introduction if the rank of the Hessian matrix for discrete systems is $(n-r); 0<r<n$, the systems can be treated as field systems. It can be observed that the treatment of Lagrangians as field systems is always in exact agreement with the general approach. The equations of motion (1.4) are equivalent to the equations of motion (1.9). Besides, the constraints (1.6) are equivalent to the equations (1.10). The consistent theory in the treatment of Lagrangians as field systems also leads to two types of constraints: a $B$--type which contains at least one member of the set $\left\{\dot q_\mu ,{\partial q_a\over\partial t},{\partial q_a\over\partial q_\mu}\right\}$, and an $A$--type which contains coordinates only. As we have seen, in the first example $F_1$ and $F_2$ are $B$--types; while the constraint $F_1$ in the second example is an $A$--type. In the general approach the constraints can be obtained from the Euler--Lagrange equations, whereas, in the treatment of Lagrangians as field systems, the constraints can be determined from the relations (1.5,6) and the new constraints can be obtained using the variations of these relations. \vspace{1cm} \noindent{\bf Acknowledgments} The author would like to thank the International Centre for Theoretical Physics, Trieste, for hospitality. \newpage
\section{Introduction} Magnetic ordering phenomena of both classical and quantum antiferromagnets on non--bipartite lattices are a fascinating subject. The simplest and most frequently studied model of this type is the Heisenberg antiferromagnet on a triangular lattice (HAFT). While the question of whether or not the typical features of this model have been observed in experiments is still a controversial issue \cite{EXP90}, the theoretical understanding of these features has advanced rapidly during recent years \cite{ADJ90,ADM92,AWE92,LR93,DE93}. In contrast to antiferromagnets on bipartite lattices, the HAFT exhibits non--collinear magnetic order in its classical and most likely also in its quantum ground state. As a consequence, the order parameter of the HAFT is represented locally by a set of three mutually orthogonal unit vectors or, alternatively, by a rotation matrix which defines the local orientation of this set relative to some fixed frame of reference. Renormalization group (RG) studies of appropriate nonlinear sigma (NL$\sigma$) models \cite{ADM92,AWE92} have revealed a number of interesting properties of the HAFT. The symmetry of the model was found to be dynamically enhanced from $O(3)\otimes O(2)$ to $O(4)$, and in a two loop RG calculation for the classical HAFT \cite{ADM92}, the temperature dependence of the correlation length $\xi$ was obtained as \begin{equation} \xi = \Delta \; C^{\xi}_{RG} \; \sqrt{T/B} \; e^{B/T}\; , \label{1} \end{equation} where $\Delta$ is the lattice constant, $B=\sqrt{3} \pi (\frac{\pi}{4} + \frac{1}{2}) =6.994$. The prefactor $C^{\xi}_{RG}$ is left undetermined by the RG calculation. It follows from topological considerations that the order--parameter field of the HAFT allows for excitations of the form of $Z_2$ vortices \cite{KM84,M79}. A numerical study of the classical HAFT \cite{KM84} has revealed that these vortices become abundant above a threshold temperature $T_{th} \simeq 0.3$ and that they unbind for $T>T_{th}$, similarly as the $Z$ vortices in the planar $XY$ model above the Kosterlitz--Thouless (KT) transition temperature $T_{KT}$ \cite{KT73}. Further evidence for this similarity beween the dissociation mechanism of the $Z_2$ vortices and that of the $Z$ vortices has been provided by Kawamura and Kikuchi \cite{KK93}. In recent work \cite{WEA94}, we have studied the influence of the vortices on the partition function of the classical HAFT on the basis of the NL$\sigma$ model. While a true KT--type phase--transition can be ruled out for the HAFT, our results suggest that for $T>T_{th}$ the vortices will affect the properties of the HAFT rather drastically. In particular, the disorder induced by the unbinding of the vortices can be expected to lead to a crossover from the T--dependence Eq. (\ref{1}) of the correlation length in the low temperature regime $T<T_{th}$ to a KT--type behavior \begin{equation} \xi = \Delta \; C^{\xi}_{KT} \; \exp( b/(T-T_{th})^{\frac{1}{2}} ) \label{2} \end{equation} in the high temperature regime. It is the aim of the present paper to supplement our recent analytical study \cite{WEA94}, which was based on a continuum description of the HAFT by a Monte Carlo (MC) simulation of the original lattice model. A similar study has recently been published by Southern and Young \cite{SY93}. While we closely follow the method of these authors, our conclusions will be quite different from theirs. In the next Section, we first give a brief account of the technique we used. Subsequently we present the results for the correlation length $\xi$ and the antiferromagnetic susceptibility $\chi({\bf Q})$. While these quantities are directly accessible to simulations in the high temperature regime, where the disordering effect of the vortices limits the range of the correlations, the key quantity to be computed in the low temperature regime is the spin stiffness \cite{CADM94,SY93}. Our numerical results for this last quantity will be presented and discussed in Section III. Finally, we summarize the evidence for a vortex induced crossover transition in Section IV. \section{Correlation Length and antiferromagnetic Susceptibility} The classsical Hamiltonian of the triangular Heisenberg antiferromagnet can be defined as \begin{equation} H = \sum_{<i,j>} {\bf S}_i \cdot {\bf S}_j \;\;. \label{3} \end{equation} Here, the ${\bf S}_i$ are three dimensional unit vectors and the sum extends over all distinct pairs of nearest neighbor sites of a triangular lattice of $L^{2}$ sites. The exchange constant has been set to unity. The classical ground state of the Hamiltonian Eq. (\ref{3}) is a coplanar arrangement in which the spins on the three sublattices are oriented at $120^{o}$ relative to each other, \begin{equation} {\bf S}_i = {\bf e}_{1} cos( {\bf Q} {\bf R}_{i} ) + {\bf e}_{2} sin( {\bf Q} {\bf R}_{i} )\;\;. \label{4} \end{equation} Here, ${\bf e}_{1}$, ${\bf e}_{2}$ are a pair of mutually orthogonal unit vectors and ${\bf Q}$ can be any one of the six vectors pointing towards the corners of the hexagonal Brillouin zone of the triangular lattice, e.g. ${\bf Q} = \frac{2\pi}{\Delta} (\frac{2}{3},0)$. The correlation length $\xi$ can be obtained assuming a Ornstein Zernicke form for the structure factor \begin{eqnarray} S({\bf q}) &=& \frac{1}{L^{2}} \sum_{i,j} e^{i {\bf q} ({\bf R}_{i} - {\bf R}_{j}) } < {\bf S}_i \cdot {\bf S}_j > \label{5}\\ &=& \frac{S({\bf Q})} {1+ \xi^{2} ({\bf q}-{\bf Q})^{2} }\;\;. \nonumber \end{eqnarray} $\chi({\bf Q})=S({\bf Q})/T$ is then the susceptibility of the system at the ordering wavevector. The spin correlations $< {\bf S}_i \cdot {\bf S}_j >$ in Eq. (\ref{5}) can be determined by MC techniques. We used the local algorithm described by Kawamura and Miyashita \cite{KM84}. The lattice is divided into independent sublattices. Then, the spins of each of these sublattices are updated sequentially. For a given spin, a new direction is chosen at random and the standard Metropolis rule is used to decide whether the new direction is to be accepted. If it is discarded, a precessional motion through a randomly chosen angle about the direction of the local mean field is performed. We apply this method to systems of linear sizes $L=12 \cdot 2^{n}, n=0,1,..5$. For the smaller systems, $n\le3$, we discard the first $2\cdot10^4$ sweeps for equilibration and average over the next $2\cdot 10^5$ sweeps. For $n=4$, we average over $4\cdot 10^5$ sweeps after discarding the initial $10^5$ sweeps, and for $n=5$, the average is over $1.8\cdot 10^6$, and $2\cdot 10^5$ sweeps are discarded. A selection of results for the correlation length $\xi$ and for the antiferromagnetic susceptibility $\chi({\bf Q})$ which have been obtained by averaging over 3-5 independent runs of these lengths is tabulated in Table \ref{tab_ksi_chi}. As will become apparent shortly, the data shown in this Table are crucial in checks of theoretical predictions for the temperature dependence of $\xi$ and $\chi({\bf Q})$. To exhibit possible finite size effects, the Table contains two pairs of data for each temperature which correspond to two different system sizes $L$, $2L$. In general, $\xi$ and $\chi({\bf Q})$ decrease with $T$, but increase with the system size $L$. If, for a given temperature $T_0$ and a given system size $L_0$, the data for $\xi$ and $\chi({\bf Q})$ exhibit no size dependence upon doubling the system size, then one can conclude that the system size $L_0$ suffices to obtain size independent data for all $T>T_0$. Data which are size independent by this criterion are marked by an asterisk in Table \ref{tab_ksi_chi}. Obviously, we cannot exclude that the data for the lowest temperature $T=0.3$ obtained for the $L=384$ system are still size dependent. Certainly, however, our data for $T=0.3$ are lower bounds to the thermodynamic limits of $\xi$ and $\chi({\bf Q})$ at this temperature. To facilitate the comparison of our data with the RG predictions we include in Table \ref{tab_ksi_chi} the values for $\xi$ and $\chi({\bf Q})$ which result from fits of the expressions Eq. (\ref{1}) and Eq. (\ref{6}) to these data \cite{SY93}. In Figs. \ref{ksi-T} and \ref{chi-T}, we show our complete sets of results for the correlation length and for the antiferromagnetic susceptibility as functions of the temperature. Obviously, for any given system size $L$, there is an inflection point in the sequences of data for $\xi$ and $\chi({\bf Q})$. This point defines a temperature $T_L$ below which both $\xi$ and $\chi({\bf Q})$ begin to exhibit finite size effects. In fact, as can be seen in Fig. \ref{ksi-T}, the correlation length increases linearly with the system size for sufficiently low temperatures $T \ll T_L$. Figs. \ref{ksi-T} and \ref{chi-T} also contain fits of different theoretical predictions to the numerical data. The dashed lines represent fits of the RG result, Eq. (\ref{1}), to our data for $\xi$ and of the form \begin{equation} \chi({\bf Q})=C^{\chi}_{RG}(T/B)^4\exp(2B/T) \;, \label{6} \end{equation} proposed by Southern and Young \cite{SY93} on the basis of RG calculations, to our data for $\chi({\bf Q})$. In these RG predictions, the constants $C^{\xi}_{RG}$ and $C^{\chi}_{RG}$ are the only undetermined parameters. In our fits, we neglect the data points for temperatures $T \leq T_L$ which contain finite size effects. In agreement with Southern and Young \cite{SY93} we find $C^{\xi}_{RG} \simeq 3 \cdot 10^{-8}$ and $C^{\chi}_{RG}\simeq 6 \cdot 10^{-12}$. The solid lines represent fits of the KT forms Eq. (\ref{2}) and \begin{equation} \chi({\bf Q})=C^{\chi}_{KT} \exp( \frac{7}{4} \cdot b/(T-T_{th})^{\frac{1}{2}} ) \label{7} \end{equation} to the data. With the KT form for $\xi$, Eq. (\ref{2}), the last expression follows from the general relation $S({\bf Q})\sim \xi^{2-\eta}$ for the structure factor at the ordering wave vector, if $\eta$ is assumed to take the value $\eta=1/4$ as for a proper KT transition. For the threshold temperature, we use the value $T_{th}=0.28$, which we can be inferred from the temperature dependence of the spin stiffness as will be discussed in the next section. This leaves the constant $b$ which is common to the expressions Eq. (\ref{2}) and Eq. (\ref{7}) and the constant factors $C^{\xi}_{KT}$ and $C^{\chi}_{KT}$ of Eq. (\ref{2}) and Eq. (\ref{7}) as fit parameters. As before, we ignore data points for $T \leq T_L$ in our fits. The results are $b=0.77$, $C^{\xi}_{KT}=0.47$ and $C^{\chi}_{KT}=0.40$. Obviously, for temperatures $T\geq0.3$, the $\exp(b/(T-T_{th})^{\frac{1}{2}})$-temperature dependence of the KT forms fits the data better than the $\exp(B/T)$-temperature dependence predicted by the RG analysis. In Figs. \ref{ksi-wT} and \ref{chi-wT} we plot $\xi$ and $\chi({\bf Q})$ logarithmically against $(T-T_{th})^{-1/2}$ so that the KT forms Eq. (\ref{2}) and Eq. (\ref{7}) appear as straight lines. These lines are seen to fit the data quite well in the temperature interval $0.30 \leq T \leq 0.34$, whereas the agreement between the curves representing the RG forms is restricted to a narrow interval around $T \approx 0.31$. In particular, we emphasize that for $T=0.3$, the RG predictions are incompatible with the data points for $\xi$ and $\chi({\bf Q})$ which are listed in Table \ref{tab_ksi_chi} with their respective errors. In this context, we recall that if our data for $T=0.3$ do not represent the thermodynamic limits of $\xi$ and $\chi({\bf Q})$, they are certainly lower bounds to these limits. Hence, the discrepancy between the RG predictions and the true values of $\xi$ and $\chi({\bf Q})$ may even be larger than has been inferred here. We note that the fit of the KT form, Eq. (\ref{7}), to the data for the susceptibility $\chi({\bf Q})$ is better than that of the RG form, Eq. (\ref{2}), to $\xi$. This may be attributable to the lower quality of the data for $\xi$ which are obtained indirectly from the Ornstein-Zernicke expression Eq. (\ref{5}) in the limit $\xi|{\bf Q}- {\bf q}|\ll 1$. In summary, we observe that our results combined with the earlier findings of Kawamura and Miyashita \cite{KM84} support the view that in the temperature range $T>T_{th}$ the spin correlations of the HAFT are decisively influenced by unbound vortices, so that a perturbative treatment of the model is inadequate in this temperature regime. It should be obvious, however, that in the above analysis of our numerical results we have been guided by our previous prediction \cite{WEA94} that, in the case of the HAFT, the vortex unbinding mechnism leads to a KT-type temperature dependence of the correlation length above a crossover temperature $T_{th}$. While we do not claim to have found compelling evidence for this prediction, we regard our numerical results as strong support for it. \section{spin stiffness} To further corroborate the above view and in order to get insight into the low temperature regime, where the correlation length exceeds the accessible system sizes, we also determined the spin stiffness $\rho$ in our simulations. The diagonal components $\rho_{\alpha}, \alpha = 1,2,3$, of the spin stiffness tensor are the second derivatives of the free energy density $f(\theta_{\alpha})$ with respect to the twist angles $\theta_{\alpha}$ of the spins around three mutually orthogonal axes ${\bf e}_{\alpha}$ \cite{CADM94,SY93}, \begin{eqnarray} \rho_{\alpha} = &-& \frac{2}{\sqrt{3} L^2} < \sum_{<i,j>} \left[ {\bf S}_i \cdot {\bf S}_j - ({\bf S}_i \cdot {\bf e}_{\alpha}) ({\bf S}_j \cdot {\bf e}_{\alpha}) \right] ({\bf u} \cdot {\bf e}_{ij})^2 >\nonumber \\ &-& \frac{2}{\sqrt{3} T L^2} < \left[ \sum_{<i,j>} {\bf S}_i \times {\bf S}_j \cdot {\bf e}_{\alpha} ({\bf u} \cdot {\bf e}_{ij}) \right]^2 > \;\; . \label{8} \end{eqnarray} Here, ${\bf u}$ is the lattice direction along which the twist is applied and ${\bf e}_{ij}$ is the direction of the bond between nearest neighbor lattice sites $i$ and $j$. The prefactor in Eq. (\ref{8}) has been chosen such that Eq. (\ref{8}) is the stiffness per unit area. In the simulation, the thermal averages on the right hand side of Eq. (\ref{8}) are replaced by averages over configurations which are generated by a large number of successive MC sweeps. For a finite system, the ordered spin structure will change its orientation in space in the course of the simulation. In a sufficiently large number of sweeps, one will therefore measure the average spin stiffness \begin{equation} \rho = \frac{1}{3} \sum_{\alpha=1}^{3} \rho_{\alpha} \;\;. \label{9} \end{equation} Since there is no long range order in the HAFT for any finite temperature, $\rho$ must vanish for the HAFT in the thermodynamic limit for any finite temperature. However, for finite system sizes $L$, $\rho$ will be finite for sufficiently low temperatures such that $\xi >L$. According to Eq. (\ref{1}), this condition should be satisfied for system sizes $L<10^3$ up to temperatures of the order of unity, unless the constant $C^{\xi}_{RG}$ is exceedingly small as has been suggested by Southern and Young \cite{SY93}. As we have argued above, the results for $\xi $ in the high temperature regime, $T>T_{th}$, should not be interpreted as evidence for such a small value of $C^{\xi}_{RG}$. In fact, a naive integration of the 2--loop renormalization group equations which starts with the microscopic parameters of the HAFT as initial values yields the result Eq. (\ref{1}) with $C^{\xi}_{RG}=\sqrt{(\frac{\pi}{4}+\frac{1}{2})} e^{- (\frac{\pi}{4}+\frac{1}{2})} = 0.314$. In our simulations, we determine the three $\rho_{\alpha}$ separately in each sweep and thus obtain the averages $\rho$ for each sweep. In Fig. \ref{rho-T}, we show the average spin stiffness as a function of the temperature for system sizes $L=12,24, \dots 384$. The steep drop in $\rho$ which occurs as $T$ increases beyond $T_{th} \simeq 0.28$ is consistent with the rapid decrease of $\xi$ in the same temperature regime in which the vortices become unbound \cite{KM84}. In contrast to the spin stiffness of the planar XY model, $\rho$ does not saturate in the low temperature regime $T<T_{th}$ with increasing system size $L$ but decreases with increasing $L$. This behavior is to be expected, since in contrast with the correlation length of the XY model, the correlation length of the HAFT remains finite for low temperatures, where the vortices are bound in pairs. As $T \rightarrow 0$, our data for $\rho$ approach the correct limiting value $1/\sqrt3$. In Fig. \ref{rho-L}, we display $\rho$ for various temperatures and system sizes. The data are averages over 3-5 independent runs of lengths comparable to those which we have described above, error bars indicate the standard deviation. In the low temperature regime, the thermodynamics of the classical HAFT should be captured by the appropriate NL$\sigma$ model \cite{DR89,ADM92,AWE92}. On the basis of a renormalization group treatment of this model, Azaria {\it et al.} \cite{ADJM92} have made detailed predictions for the dependence of the spin stiffness on the linear system size $L$ and the correlation length $\xi$. In their MC study of the classical HAFT, Southern and Young \cite{SY93} found excellent agreement with the predicted $L$ dependence of the spin stiffness tensor at the temperature $T=0.2$. In order to be able to compare our numerical results with the predictions of the RG analysis of the NL$\sigma$ model, we integrated the two loop RG equations of Azaria {\it et al.} \cite{ADJM92} starting from the initial conditions $\rho (L=\Delta)=1/\sqrt3$ and $\rho _3(L=\Delta)/\rho _1(L=\Delta)=2$. Here, $\rho _1$ and $\rho _3$ are the two main components of the spin stiffness tensor with respect to the reference frame of the local order parameter \cite{SY93}. By the above initial conditions we identify $\rho _1$ and $\rho _3$ with their microscopic values on the scale of the lattice constant $\Delta$. We find that in the temperature regime under consideration, $T\leq 0.3$, the average stiffness $\rho$ varies linearly with $\ln L$ to a very good approximation on the scale $12\leq L\leq 384$ covered by our simulations, see Fig. \ref{rho-lnksi} below. {}From the RG equations, one can also infer that the slope $\rho '(L)=-d\rho (L)/d\ln L$ decreases from $\rho '=T/(3\pi)$ to $\rho '=T/(4\pi)$, when $L$ increases from a value of the order of the lattice constant, $L \sim \Delta$, to a value of the order of the correlation length, $L \sim \xi $. In Fig. \ref{rho-L}, the straight lines are least squares fits to the data. The slopes of these lines, normalized to the maximal theoretical slope $\rho '=T/(3\pi)$, are tabulated in Table \ref{Gefaelle}. For $T\leq 0.25$, the slopes are seen to be rather close to their maximal value which obtains, when $L$ is of the order of the lattice constant. This is not unexpected since according to the RG calculations, the correlation length is many orders of magnitude larger than our maximal system size $L=384$ for these temperatures. The larger values of $\rho '$ which we find for $T\geq 0.28$ are incompatible with the RG theory. Hence we conclude that for $T\geq 0.28$ unbound vortices, not being taken into account by the RG analysis, begin to limit the range of the spin correlations in the HAFT. If one defines the correlation length $\xi$ through the matching condition $\rho (L=\xi)=0$, then the result of the integration of the two loop RG equations can be cast into the following form \begin{equation} \rho = f(T,L) \; \ln (\xi/L) \;\; . \label{10} \end{equation} This is shown for two different temperatures in the two graphs in Fig. \ref{rho-lnksi}. Obviously, the function $f(T,L)$ depends weakly on L. The relation Eq. (\ref{10}) makes it possible to obtain the correlation length from the Monte Carlo simulation, even in the low temperature regime where $\xi$ is much larger than the system size $L$. Inserting our MC data for $\rho $ into Eq. (\ref{10}) and solving for $\xi$, we obtain the data points shown in Fig. \ref{Reklame} for $T \leq 0.28$. This Figure also includes the data for $T \geq 0.28$ which have already been shown in Fig. \ref{ksi-T}. The dashed and solid lines in Fig. \ref{Reklame} represent fits of the RG and KT forms, Eqs. (1) and (2), to the MC data for $T \leq 0.28$ and $T \geq 0.28$, respectively. \section{Summary} We have shown that the MC simulation of the HAFT yields compelling evidence for the influence of vortices on the spin correlations and hence on the thermodynamics of this model. In agreement with earlier findings by Kawamura and Miyashita \cite{KM84} we find that the disordering effect of the vortices sets in rather abruptly at a temperature $T_{th} \simeq 0.28$. Up to this temperature, our data for the spin stiffness $\rho$ and the ensuing temperature dependence of the correlation length $\xi$ are consistent with the predictions of the RG analysis of the HAFT which ignores the existence of topological defects such as vortices. For $T>T_{th}$, however, the simulation yields temperature dependences of the correlation length $\xi$ and the antiferromagnetic susceptibility $\chi({\bf Q})$ which are incompatible with the RG predictions. Instead, in this temperature regime, the temperature dependences of $\xi$ and $\chi({\bf Q})$ which follow from the KT picture of unbinding vortex pairs provide satisfactory fits of the data. A rapid increase of the density of unbound vortices for $T>0.3$ had already been found by Kawamura and Miyashita in their simulation of the HAFT \cite{KM84}. It had not been clear, however, whether this phenomenon would lead to the same temperature dependences of the correlations of the HAFT as had been predicted for the planar XY model by Kosterlitz and Thouless \cite{KT73}. In our recent analytical study \cite{WEA94}, we were led to the conclusion that this should indeed be the case. The present numerical study fully supports this conclusion. As we have discussed in Ref. \cite{WEA94}, the crossover transition from the RG type behavior to the KT type behavior of the correlations of the HAFT cannot imply a phase transition in the proper sense, because the correlation length is finite both below and above the transition temperature $T_{th}$. The results shown in Fig. \ref{Reklame} indicate, however, that the transition happens in a very narrow interval around $T_{th}$. Therefore we consider it possible that the derivative of the correlation length with respect to the temperature develops a discontinuity at $T_{th}$ in the thermodynamic limit. \subsection*{Acknowledgement} The numerical calculations were carried out in part at the Regionales Rechenzentrum Niedersachsen, Hannover.
\section{Introduction:} The CCT concept has been outlined in a previous publication \cite{nim94} and described in detail in an internal CLEO collaboration document\cite{cbx}. The technique uses the correlation between photon pathlength and $\check{\rm C}$erenkov production angle to infer this angle by measuring the time taken for the totally internally reflected $\check{\rm C}$erenkov photons to ``bounce'' to the end of the radiator. The most typical radiator geometry is that of a quartz bar having rectangular cross-section, typically a few centimeters on a side, and having a length of about a meter. This is very similar to the radiators in the DIRC design of Ratcliff\cite{dirc}. The key parameters of a CCT system will all be related to timing performance. In practical terms this translates to photo detector efficiency and transit-time jitter, quality of the radiator bars in terms of geometry and transparency, and our ability to make high quality optical couplings between bars and detectors. All of these issues are addressed below. \section{Timing Considerations:} The technique relies exclusively on fast timing information, hence an important aspect of its evaluation will be to study the various available photon detector technologies. The most obvious of these, and hence the initial choice, are photomultipliers. Most CCT geometries under study involve radiators whose cross-section have widths of about 4 cm, and thicknesses varying between 1 and 4 cm. A standard 2" photomultiplier provides an acceptable match to this. For example, the 12 stage Hamamatsu H2431 has an active photocathode diameter of 4.6cm, which will cover 93\% (100\%) of the end of a $4 \times 4$ cm$^2$ ($4 \times 2$ cm$^2$) bar. This particular tube, which was used extensively in our beam-test experiments\cite{kichimi}, is also one of the fastest available 2" photomultiplier tubes, having a single photoelectron timing jitter ($\sigma_{1pe}$) of about 160ps. Even this seemingly outstanding performance may become a limiting factor in a real CCT device. For a relativistic charged particle normally incident on a quartz bar (n=1.4) at a distance $z=1$ meter from its end, the angular resolution of the $\check{\rm C}$erenkov angle per unit of timing uncertainty ($d\theta/dt$) is about 150 mr/ns, degrading to about 550 mr/ns for tracks incident at 25 degrees from normal incidence\cite{zdepencence}. If we set 15 mr resolution as a conservative goal (this would provide $>2\sigma~ K-\pi$ separation up to about 2 GeV/c momentum using CCT alone\cite{addtof}) we see that we will need timing resolution of 100 ps (27 ps) at $0^\circ$ ($25^\circ$) incidence. It is clear that if we use conventional photomultipliers we will have to rely on photon statistics to obtain the required timing performance. We would expect the resolution to scale roughly as $\sigma_{1pe}/\sqrt{N_{prompt}}$, where $N_{prompt}$ is the number of photoelectrons detected in a time window starting at the nominal photon arrival time, having a width small compared to $\sigma_{1pe}$\cite{prompt}. Clearly, both phototube jitter and photon yield are critical parameters. We have measured the transit-time spread of the 1", 10 stage Hamamatsu H5321 photomultiplier, the fastest conventional tube commercially available at the time of our test. Our experimental setup was similar to that used by Kuhlen\cite{kuhlen}, and is not described here. We measure the standard deviation of the single-photon transit time to be $\sigma_{1pe} = 82\pm 25$ ps, consistent with Hamamatsu's claim of 160 ps FWHM. The timing distribution is observed to have a tail on the high-time side, consistent with expectation \cite{kuhlen}. Although the performance of both the 1" tube tested in our laboratory and the 2" tube used in our beam test were sufficient for the CCT prototypes built so far, the availability of a photon detector with much smaller jitter would let us trade timing performance for photon yield. This in turn would make thinner radiators a possibility, or perhaps even the use of UV plastic as a replacement for quartz. Photon yield is discussed in the following section. \section{Optical Transmission Studies:} The number of $\check{\rm C}$erenkov photons produced by a charged particle traversing a radiator depends only on the speed of the particle, and the index of refraction of the material, which is fairly constant at around $n = 1.47$ for most solid radiators. The parameters that will ultimately determine the number of detected photoelectrons will therefore be those that describe the transport of photons from their production point to the photodetectors. Most notable of these is the optical transparency of the radiator material and any coupling compounds used. In this section we describe measurements of transmission spectra performed on three possible radiator materials (fused silica quartz and two kinds of plastic), as well as several optical coupling compounds. Radiation length measurements are presented in another paper\cite{kichimi}. The transmission vs. wavelength studies were performed using McPherson 218 0.3m scanning monochromator illuminated by a deuterium lamp. The sample under test was mounted at the exit port of the monochromator, and the transmitted light intensity was measured by a photomultiplier tube feeding its anode current to a precision electrometer. The wavelength scan and data acquisition was controlled by a PC with a custom-built interface. Anode-current versus wavelength spectra were accumulated for each of the materials under study, and each was normalized to a ``no sample'' spectrum to yield the transmission curves presented here. In each case we look for the wavelength at which the material transmission ``turns on'', (defined as the 50\% transmission point), and the width of the turn-on, (defined as the change in wavelength between 10\% and 90\% transmission). Three radiator materials were tested: (i) Nippon fused silica quartz machined into a $4\times 4\times 50$~cm$^3$ bar by Surface Finishes Inc, (ii) A $2\times 2\times 0.5$~cm$^3$ piece of Mitsubishi Acrylite 000 provided by KEK, and (iii) A $2\times 2\times 0.5$~cm$^3$ piece of Kuraray plastic LightGuide-S provided by KEK. The transmission spectra of these materials is shown in Figure \ref{transmission}. The quartz is clearly superior, turning on at a lower wavelength with a sharper edge than either of the organic materials. \begin{figure} \centerline{\psfig{file=fig1.ps,bbllx=92bp,bblly=176bp,bburx=530bp,bbury=626bp,height=5cm}} \caption{Transmission spectra for quartz (solid) acrylic (dashed) and plastic (dotted) radiators.} \label{transmission} \end{figure} \smallskip We also tested four possible optical coupling compounds: (i) Oken 6262A grease\cite{grease}, provided by KEK, (ii) Dow Corning Q2-3067 grease, (iii) General Electric Viscasil grease, and (iv) Dupont Krytox oil. The results are summarized in Table 1. \begin{table} \begin{center} \caption{Transmission properties of three possible CCT radiators and four optical coupling compounds} \vskip 2 mm \begin{tabular}{|c|c|c|} \hline Material & $\lambda_{50\%}$ (nm) & $\Delta\lambda_{10-90}$ (nm)\\ \hline Quartz & $265\pm 3$ & $15\pm 3$ \\ Acrylic & $290\pm 5$ & $25\pm 5$ \\ Plastic & $285\pm 5$ & $35\pm 10$\\ \hline \multicolumn{1}{|c}{Coupling Compound} & \multicolumn{2}{c|}{}\\ \hline Oken & $280\pm 5$ & $25\pm 5$ \\ Corning & $275\pm 5$ & $20\pm 5$ \\ Viscasil & $195\pm 10$ & $25\pm 10$ \\ Krytox & $< 150$ & -- \\ \hline \end{tabular} \end{center} \label{transprop} \end{table} The Oken and Corning compounds are quite similar in performance, whereas the Viscasil has a significantly lower wavelength cutoff. Remarkably, the cutoff of the Krytox oil was below the reach of our apparatus. This material is a fluorinated lubricant developed by Dupont, and has other interesting properties\cite{dupont}. \section{Quartz Quality Studies:} The performance of a CCT detector is determined by the time resolution of the photon detector and by the quality of the quartz radiator. These two parameters will also set the total cost of the detector. The price of the quartz is controlled by its UV transparency and by the quality of the surface finish. The second point is closely related to the geometry of the bar. Standard polishing machines limit the bars to lengths less than approximately 1.20 m and square bars are easier to handle than designs with a large aspect ratio. Parameters relevant for a CCT device include the transparency of the quartz, losses due to absorption or imperfections and the surface quality needed to preserve the angle information. We have devised a series of tests to study these parameters and the results of our investigation are presented in the following sections. Only two manufactures were able to produce quartz bars to our specification: Surface Finishes Inc. (SF) and Zygo Inc. We bought two 50 cm long bars from SF and one 120 cm long bar from Zygo in order to find out if material of sufficient quality can be obtained. Details of the specification can be found in Table 2. \begin{table} \caption{Quartz Bar Specifications} \vskip 2 mm \begin{center} \begin{tabular}{|c|c|c|} \hline & SF & Zygo \\ \hline Grade & \multicolumn{2}{|c|}{Semiconductor}\\ Length & 0.5 m & 1.2 m\\ Width & 4.0 cm & 4.0 cm \\ Height & 4.0 cm & 2.0 cm\\ Roughness & & 7 \AA (rms) \\ Flatness & 0.002 mm & \\ Parallelism & 0.02 mm & 0.01\\ Perpendicularity & 90$^o\pm$30''& \\ Edges & beveled & sharp \\ \hline \end{tabular} \end{center} \label{spec} \end{table} UV-grade quartz which is transparent down to wavelengths around 150 nm should be the material of choice for a {\v C}erenkov detector but our simulation shows that the gain in the number of photons is offset by the large dispersion in this wavelength region. For this reason and to reduce the overall costs we have selected semiconductor grade quartz. In a CCT detector the {\v C}erenkov photons propagate toward the photon detector via total internal reflection. During this process a small amount of energy travels a short distance as a surface wave outside the medium. Ideally this is loss free but any surface imperfection will change this. With the large number of total internal reflections possible in a CCT device it is important to study this experimentally. We have set up a HeNe laser with several Al mirrors that give us complete control over the light direction in the quartz bar. Photodiodes are used to measure the intensity of the incident, reflected and transmitted laser beam. By changing the angle of incidence we varied the number of total internal reflections and obtained the results shown in Figure \ref{reflection}. Fitting to a straight line gives $0.4 \pm 0.07$\% loss per reflection for the bar from Surface Finishes and for the Zygo bar we find a $0.1\pm 0.03$\% loss per reflection. For 50 bounces this results in a total loss of 5\% (20\%) for Zygo (SF). \begin{figure} \centerline{\psfig{file=fig2.ps,bbllx=92bp,bblly=176bp,bburx=530bp,bbury=626bp,height=5cm}} \caption{Loss as function of the number of internal reflections for the Surface Finishes bar (a) and the Zygo bar (b).} \label{reflection} \end{figure} In our study we have carefully avoided the corners of the bar. Should the light hit a corner, losses are significantly larger: For the Zygo bar we measured 4\% per reflection, and losses are even larger for the bar from Surface Finishes due to the beveled edges. Semiconductor grade quartz can have some inclusions such as air bubbles. As much as 10\% of the light was lost when we directed the laser beam on a small air bubble in the Zygo bar. At a momentum of 2.5 GeV or higher the pion and kaon {\v C}erenkov angles differ by only a few mr and this small angle difference has to be preserved as the {\v C}erenkov photons travel toward the photon detector. This places severe requirements on the geometry of the quartz bar. In particular, we were interested in determining if the manufacturers could achieve the parallelism and perpendicularity specified. In a first series of tests we used an optical bench to verify that the macroscopic (geometrical) properties of both quartz bars were within our specification. We then searched for small scale variations in the thickness that could change the photon angles. Using two lenses we widened the beam of a HeNe laser to a spot size of approximately 1 cm which we directed on one of the long sides of the quartz bar. Changes in the interference pattern between light reflected at the near and far side reveal variations in the thickness of the quartz bar. \begin{figure} \centerline{\psfig{file=fig3.eps,bbllx=42bp,bblly=32bp,bburx=513bp,bbury=434bp,height=7cm}} \caption{Thickness variation over the length of the bar from Surface Finishes (a). For the Zygo bar both height and width variations are shown (b).} \label{interference} \end{figure} The results of this study are shown in Figures \ref{interference}a and b for the Surface Finishes bar and the Zygo bar, respectively. The distance between two interference fringes corresponds to a thickness variation of $\Delta s \; = \; \lambda/2n \; = \; 0.2 \mu m$\cite{factoroftwo}. While the wide sides of the Zygo bar are very flat and parallel, significant curvature is found for the narrow sides confirming the initial expectation that a large aspect ratio makes the polishing process more difficult. For both bars, thickness variations become more pronounced close to the corners. Quantitatively, we find a thickness variation of approximately 1-2~$\mu$m for the Surface Finishes bar and better than 0.5~$\mu$m (2-3~$\mu$m, narrow sides) for the Zygo bar. Both values are acceptable for a CCT detector. We gratefully acknowledge the support of the Department of Energy, the National Science Foundation, and the A. P. Sloan Foundation. We would also like to express our thanks to Prof. S.Iwata and Prof. F.Takasaki at KEK for their support of this work.
\section{History} Remote observing has been widely used in radio astronomy for the last decade. However, in the optical and infrared domain very few observatories have been successful in supporting it. One reason for this difference is that optical/infrared telescopes are more demanding in the sense of object acquisition and amount of science data produced. Another reason could be that optical astronomers have less trust in the performance of the telescope and instruments. In the early 1980's a number of observatories started experimenting in remote observing (Raffi \& Tarenghi 1984, Longair et al. 1986, Raffi \& Ziebell 1986). Most of these early attempts failed, mainly because the technology was not yet matured. The communication links were not reliable and the data transmission rates were too low. It also became evident that it was very difficult to operate a telescope originally designed for local control in remote mode. In the late eighties the situation became more favourable. First, there was a general improvement in communication infrastructure: cheaper, more reliable and faster, not to forget the development of Internet and, more recently, of the World-Wide Web. Second, a new generation of 3--4-m class telescopes was designed and went into operation. For some of these, remote observing had been foreseen already during the design phase (Loewenstein \& York 1986, Raffi et al. 1990). This led to the successful use of several different modes of remote observing, ranging from fully automatic telescopes to interactive long-distance instrument and/or telescope control. In 1992 a workshop dedicated to remote observing was held in Tucson (ed. Emerson \& Clowes 1993), where many of the available forms were extensively discussed. We will first summarize the terminology adopted at this conference for the various remote-observing modes, before discussing the pros and cons and the technical requirements for remote observing. In the second part of the paper we discuss the experience obtained at ESO, with emphasis on the observing efficiency, the support structure and the feedback from astronomers. \section{Definitions of Remote-observing Modes} \hskip \the\parindent {\it Robotic Telescopes} - A telescope and instrument which is programmed beforehand and needs little or no interaction during the night. An entire night's observing program may be downloaded during the day and the scientific data uploaded during the observations or later. {\it Remote Engineering} - A remote engineer performs engineering activities, e.g. installations, diagnostics, troubleshooting, on local equipments. Although this is not an observing mode it should be included in the context, because if remote engineering is available, remote observing may come for free. Remote engineering can be very effective, especially since often engineers need to come to the telescope for relatively minor tasks and the travel distance can be significant. {\it Service Observing} - An astronomer fully specifies the objects to be observed and the instrument configuration beforehand, and the actual observation is carried out by the observatory staff. The astronomer is awarded observational data rather than telescope time. {\it Passive Remote Observing} - A remote observer monitors an observation carried out at a telescope by a collaborator or service observer. The remote observer has access to the data obtained and, optionally, observation parameters, and interacts with the local observer via voice, ``talk'', or e-mail. This is also called {\it Eavesdropping}. It is probably fairly common, but observatories tend not to keep track of this. {\it Active Remote Observing} - A remote observer interactively controls an instrument and, optionally, a telescope at an observatory. In most cases a local telescope operator is required for safety reasons and sometimes to operate some telescope and auxiliary equipment. This is also called {\it Remote Control}. Combinations of these observing modes are of course possible, and may in fact give additional advantages. {\it Robotic telescopes} are by far the most popular of these options. They are typically small telescopes (50cm) and there are two important uses: the first is for long-term monitoring programs (an example is the network of small telescopes looking for solar oscillations) and the second is for educational purposes: some universities now have a small robotic telescope where astronomy students can send in requests electronically for short observations. At the other extreme, {\it active remote observing} is rare. The only large telescope for which this mode is the norm is the 3.5m on Apache Point Observatory (APO), which is run by a consortium of US universities. This telescope is operated for 80\% of the time available for scientific observations in remote control. Typically, a night will be split between several projects, and each astronomer can carry out the observations from his or her campus. ESO has two telescopes which can be operated in remote control: the 1.4m CAT which is used for high-resolution spectroscopy mainly of bright stars, and the 3.5m NTT. \section{Pros and Cons} The motivation for remote observing has been debated for some time in the literature and a high level of agreement on the main arguments has been obtained (e.g. Emerson \& Clowes 1993). These are summarized in Table 1 and briefly discussed below. \begin{table} \caption{Weighing the pros and cons of remote observing} \begin{tabular}{||l|l||l|l||} \hline \hline Proven & Likely & Likely & Proven \\ advantages & advantages & disadvantages & disadvantages \\ \hline \hline Flexibility & More than one & More difficult & Expensive \\ & participating & to concentrate & - personnel \\ & astronomer & on observing & - communication \\ \hline Shorter obser- & May save costs & Not as efficient& \\ ving programs & in some cases & as classical & \\ \hline Convenient for & & & \\ astronomer & & & \\ \hline \hline \end{tabular} \end{table} The main arguments in favour of remote observing are all related to the fact that the astronomer does not need to travel to the observatory. This is a strong argument if the observatory is in an inaccessible place, the most extreme example being the planned observatory on the Antarctic Plateau for which remote observing is envisaged (Burton 1995). It also allows for the observing schedule to be made at short notice (as is done at the VLA). However, flexible and queue scheduling is not recommended in active remote control and/or eavesdropping: for the astronomer to be directly involved the time of the observations needs to be known in advance. Instead it is possible to schedule shorter observing programs, in units of hours instead of nights. Monitoring programs also become easier to schedule without an astronomer at the observatory having to be involved. The elimination of long travel times and acclimatization results in savings of astronomer's time -- an important point for researchers at universities who may have teaching duties. Remote observing very often allows more than one astronomer to take part in the observations. The impression at ESO is that, for the large telescopes, astronomers prefer to come with more than one person: while one person does the observing, the other concentrates on the on-line data analysis. For remote control, financial support for an extra observer is easier to arrange. ESO also attracts some Eastern-European astronomers for remote observing who could not possibly afford to travel to Chile. For them remote observing can save costs, but for the observatory this will in general not be true unless the required communication links are free, or very cheap, and little additional support is required. Cost is the main argument against remote observing. In order not to degrade the scientific efficiency, the remote observer must get the same support as the local observer. This means that often the support personnel have to be duplicated. Sufficient bandwidth is also required in order not to have idle telescope time and the cost of this bandwidth can very easily overtake the savings in travel cost. Table 2 shows typical cost estimates for the case of ESO, illustrating that remote observing only reduces total costs when it is used for a significant fraction of the observing proposals for at least two telescopes. A second argument against remote observing is, that, due to limitations in bandwidth, there is a time delay before the observer sees the data coming from the instrument. (This problem is not unique to remote observing but is experienced by everyone observing with CCDs.) This becomes a problem whenever the extra delay is a significant fraction of the average integration time per exposure. Finally, if the observations are done from one's own office, the distractions due to the normal office activity may easily cause a loss of efficiency. \begin{table} \caption{Comparing operational cost (in kDM/month)} \begin{tabular}{||l||l||l|l|l|l||} \hline \hline No. of telescopes & 0\% RO & 10\% RO & 30\% RO & 50\% RO & 100\%RO\\ and observing runs& 0 rem op& 1 rem op& 2 rem op& 2 rem op& 3 rem op\\ \hline 1 tel.5 runs/mon. & 20 & 46 & 53 & 50 & 52 \\ 1 tel.10 runs/m. & 40 & 61 & 68 & 62 & 56 \\ 1 tel.20 runs/m. & 80 & 102 & 100 & 86 & 64 \\ CAT+NTT 20r/m & 80 & 102 & 100 & 86 & 64 \\ \hline & 0 rem op& 1 rem op& 3 rem op& 4 rem op& 6 rem op\\ \hline 4 tel.80 runs/m. & 320 & 322 & 291 & 250 & 142 \\ \hline \hline \end{tabular} {\footnotesize The following assumptions are used for the calculations:\\ Classical observing run (travel, accommodation etc.) = kDM 4\\ Remote observing run (travel, accommodation etc.) = kDM 0.8 \\ Remote operator (rem op) = kDM 10/month\\ Link cost (50\% of total cost) = kDM 18/month\\ Same local support for classical and remote observing} \end{table} \section{Requirements and Techniques} The main requirement for all modes of remote observing except robotic telescopes is a fast data transfer from the observatory to the remote site. Data files produced by modern instruments are large (a typical CCD frame is 2 to 8 Mbytes), and they should ideally be transferred in no more than a few minutes. To achieve this a fairly high bandwidth is required, although modern data compression algorithms can improve the transfer rate. Compression based on the H-wavelet transform has been successfully applied at a number of observatories. Both the availability and the reliability of the communication link need to be very high. A second requirement is that the telescope and instrument must be reliable and stable. Solving technical problems is generally more time consuming during remote observing, due to the extra feedback time. A good way to limit unforeseen problems is by keeping instrument change-overs to a minimum. Limiting the instrumentation is also a better way to ensure adequate know-how at the remote-observing site. Internet is certainly the main carrier of remote observing traffic. The tremendous advances in bandwidth and number of users and, more important, the fact that Internet is still ``free'', makes it the obvious first candidate to implement the communication link. The World-Wide Web gives a convenient interface which is already used for some robotic telescopes. A future commercialization of Internet may make some of these advantages obsolete. Also, for observatories located in remote places it may be difficult to obtain a fast Internet access. As an alternative, dedicated links to more populated areas with Internet access can be acquired. Although dedicated links are more expensive they also offer advantages like guaranteed bandwidth and propagation delays. A {\it Remote Observing Center} is a dedicated geographical site from where the observation is carried out. This is typically located in a major astronomical research institute and the remote observer has to travel to this site. The advantages are that expert knowledge, similar to a local observatory, can be built up in order to provide accurate support to the visiting astronomer. {\it Distributed Remote Observing} is when any site with a network connection, normally Internet, can perform remote observing. The obvious advantage is that no travel is required. However, it assumes that the remote observer has expert knowledge, and does not need dedicated personal support. \section{ESO Experience} ESO currently supports {\it active remote observing} with two telescopes. The 1.4m CAT has been operated routinely in active-remote-observing mode about 50\% of the time for the last six years (Baade 1993). The NTT was the first ESO telescope specifically designed to accommodate remote observing. Remote observing was first offered to the user community in 1993, and since then about 15\% (four to five nights per month) of the time is scheduled in active-remote-observing mode. Early experiences with the system are described in Balestra et al. (1992), Baade at al. (1993) and Wallander (1994). At ESO Headquarters in Garching a {\it Remote Observing Center} provides the interface for remote observers to the La Silla Observatory in Chile. Observing from other sites and institutes is not supported. A dedicated satellite link provides the bandwidth between the two sites. The system provides fully interactive {\it active remote observing} based on dedicated remote software executing at the remote site. The architecture of the NTT remote observing system was reported in detail in Wallander (1993). In November, 1994 the $64\,$kbps (used by NTT) and the analogue (used by CAT) PTT-leased communication links were replaced with a $2\,$Mbps ``roof-to-roof'' satellite link. The complete system was obtained as a turn-key project from an external contractor. Although this involved a considerable capital investment, the actual operation costs did not increase. It turned out that by leasing the bandwidth directly from the satellite provider, instead of via national PTT's, a 26-fold increase in bandwidth could be acquired for a lower price. In addition, by mounting the antennae directly on the ESO premises and becoming independent of PTT's, the reliability of the link increased. During the period November 1994 to June 1995 the availability of the link has been over 99.8\%. This should be compared to the record of the previous $64\,$kbps PTT-leased line, which in the period 1991 to 1994 had an average availability of 95\%. Figure 1 shows the data transfer rate from La Silla to Garching, averaged every five minutes, over the last ten months. The saturation of the previous $64\,$kbps link, of which 32--48$\,$kbps was allocated for data, is clearly visible. The zoom of the first two weeks of April shows that remote observing is a main, but not the only, user of the link. \begin{figure} \vspace{10cm} \caption{ Link utilization } \end{figure} \section{Observing Efficiency} We define the observing efficiency as the fraction of time between nautical twilights that the shutter of the instrument is open. This number is automatically derived from the computer-based operation log and we have routinely recorded it for the NTT since November last year. We only used data from nights without any technical problems or time lost for weather, and where the instrument, EMMI or SUSI, was used in one of the standard modes. The result is shown in Figure 2. With the exception of one remote observing run (2 nights) the obtained efficiency for classical and remote observing are very similar. Performing some statistics on the data we get the result shown in Table 3, where the first number gives the efficiency expressed as a percentage, with its standard deviation, and the number in braces gives the number of nights used in the calculation. The first night of a run is listed separately because of possible familiarization effects (``first-night syndrome''), which are indeed present. \begin{figure} \vspace{10cm} \caption{Comparing classical and remote observing efficiency } \end{figure} \begin{table} \caption{Comparing classical and remote observing efficiency} \begin{tabular} {||l|l||l|l||} \hline \hline \multicolumn{2}{|c|}{Classical observing} & \multicolumn{2}{|c|}{Remote observing} \\ First night & Later nights & First night & Later nights \\ \hline \hline 64+/-11\% (22)& 71+/-10\% (29)& 60+/-18\% (11)& 67+/-18\% (8) \\ & & 65+/-10\% (10)& 73+/-9\% (7) \\ \hline \hline \end{tabular} \end{table} Due to the still small number of nights, the difference between remote and local observing (a few per cent) can entirely be attributed to the one observing run which reached very low efficiency (which explains the larger standard deviation). Removing this run we get identical figures. We conclude from this that for normal observations remote observing is now competitive. \section{Observing Support} The support given to the remote observer requires special attention. Socially, there are more distractions making it more difficult to concentrate on observing. The time difference between La Silla and Garching is such that the second half of the night in Chile corresponds to the morning in Garching, when office life goes on as usual. Having a separate observing room is a necessity. It is also more difficult to locate people with adequate technical knowledge of the system who can introduce the observer and help with preparing the observations. This is especially true if the remote observing system is only used occasionally. It should be noted that very few people have so much observing experience on the NTT that they do not need any introduction. In practice, most astronomers only use a particular telescope a few nights per year, and they will probably try to use a different mode from the one before. In order to allow the observer efficient use of the telescope, it is essential to arrange for good support. For the NTT, we now have a group of four post-doctoral fellows and students who give introductions in Garching. During the observing night the astronomer is also supported by a remote-control operator in Garching, who controls the telescope, and a local night assistant present at the telescope. The latter is present only as a safeguard, but essential to assist in case of problems. One could describe the support problem as trying to build an observatory-environment away from the telescopes. People need to have the know-how and be aware of recent problems and changes (which requires good communication channels with the observatory). At ESO, the introduction of a ``telescope team'' for the NTT, with staff both from Garching and La Silla, has helped significantly in improving communications. The effort put into the support structure is considerable, but is necessary if remote observing is to be more than a tool for highly experienced observers, which most astronomers are not. The experience at ESO is that in almost all cases, the home institute pays for a second observer in addition to the one paid for by ESO, in the case of remote control. This 'self-support' also helps to improve the efficiency of the observations. It may be of importance from an educational point of view as well. \section{Perceptions by Astronomers} Astronomers who have used the remote control system are often, but not always, satisfied. Often their feedback has resulted in improvements of the system or of the support. When applying for observing time with the NTT, the astronomer has to specify why the program would not be suited for remote observing. We do not force people to use remote observing against their will, which would certainly backfire, but the remarks give a good impression of the perceived problems with remote observing. Occasionally, the overheads are thought to be larger in remote observing. This may be true for complicated programs, and is certainly true if the observer waits for the image to arrive before deciding what to do next, as some do. Part of this problem has been solved with the new, much faster link. More commonly, the astronomer feels that the judgment whether the conditions are photometric are more difficult in remote observing mode. It is not so clear whether this argument is correct. It is always difficult to judge how photometric conditions are when using CCDs, especially during dark time when no moon is present. The La Silla night assistant keeps an eye on the weather, but astronomers have been reluctant to accept his opinion. There have been cases where the observer wanted to continue while the moon was no longer visible. Data from the meteomonitor, which displays seeing, windspeed, temperature, etc., are displayed on-line and this is one of the most popular things to watch in the control room. Twilight flats are seen as the most demanding part of the observations, because there is little time available and the observer cannot wait for the image to arrive before knowing what the count levels were in the previous flat. The solution here is to do fast statistics on the data on the La Silla workstation, where the original file first arrives. This is perfectly possible, either by the observer working over the link or by the night assistant on La Silla who sees each image immediately after it has been read out. Good results have been obtained by using the Tyson sequences for twilight flats (Tyson \& Gal 1993). The program to calculate these sequences is available on-line. Some astronomers also express doubt about focussing. The instrument is generally focussed using one or more exposures of a random field (preferably close to the target). Here the image transfer can be very quick because it is not normally necessary to read out the full CCD. The image is normally analysed by the night assistant in La Silla, but it has been done remotely as well. This took 30 seconds longer and gave the same answer. The most serious argument used by astronomers is that their object is very faint and that it is difficult to position it at the slit. If the object is not visible on the video screen, it is necessary to take acquisition images and to move the telescope such that the selected object falls in the slit. Often two such exposures are needed. The additional overhead becomes larger if the science exposures are relatively short and there are many targets: if many acquisition images need to be taken the observer should go to La Silla. \section{Future} What is the future of remote observing? There is clearly a demand from the ESO community, seeing that a third of the optical NTT proposals request this mode. Part of this may be due to the fact that it is easier to come with more than one person in this mode. The recent experience has shown that, for many programs, remote control is competitive with local observing, being as efficient in telescope usage while giving a saving of the astronomer's time. At the same time, there is a large group of people who prefer to travel to La Silla. We will for the time being continue to offer remote observing as a service to the community, but not force it upon people. We will try to improve the system to alleviate the doubts as expressed above. A major upgrade of the NTT control system is being undertaken as part of the NTT Upgrade Project. The aim of this activity is on the one hand to verify the concept and software to be used for the VLT, on the other hand to provide an identical interface on the NTT for higher level operational tools, procedures and methods to be used on VLT. It is expected that the VLT technology and software architecture will give essential performance advantages also for remote observing. Faster computers, more efficient communication protocols, on-the-fly data compression and fast data forwarding will reduce the data-transfer rate. The limiting factor of a CCD display will become the readout time, independent of where the display unit is located. \section{Conclusions} We have shown that the observing efficiency does not degrade when using {\it active remote observing} for the ESO NTT as compared to classical observing. This allows more flexibility in scheduling, shorter observing programs, long term monitoring programs, and savings of astronomer's time. However, {\it active remote observing} is nothing else than moving classical observing to another site. It does not address the ``first-night syndrome''. To increase the scientific efficiency, service observing may be a more important observing mode than remote observing. Assuming the service observer will be at the telescope, we would expect increased demands for {\it eavesdropping} capabilities. The requirements for this to be successful are a sufficiently fast link and adequate communication facilities, i.e. not much different from those of {\it active remote observing}. The main role of {\it active remote observing} may be found in the new generation of large telescopes, where the observing runs may be very short, and for astronomers in places where travel money is difficult to get. \vskip 20pt \noindent {\bf Acknowledgements} Manfred Ziebell, with support from Joar Brynnel, has been responsible for the very successful installation and operation of the new satellite link. We thank Miguel Albrecht for providing the observing efficiency data and Jesus Rodriguez for helping us obtaining information from several observatories. The successful operation of ESO remote observing would not have been possible without dedicated support from the whole operation crew, both at La Silla and in Garching. \section{References} \begin{list} {}{\itemsep 0pt \parsep 0pt \leftmargin 3em \itemindent -3em} \item Baade D. 1993, Observing at a Distance, ed. Emerson D., Clowes R. (World Scientific, Singapore), p. 131 \item Baade D., et al. 1993, The ESO Messenger, 72, 13 \item Balestra A., et al. 1992, The ESO Messenger, 69, 1 \item Burton M. 1995, in ASP Conf. Series, 73, p559-562 \item Emerson D., Clowes R. 1993, Observing at a Distance, (World Scientific, Singapore) \item Loewenstein R.F., York D.G. 1986, SPIE 627 \item Longair M.S., Stewart J.M., Williams P.M. 1986, Q. Jl R. Astr. Soc. 27,153 \item Raffi G., Tarenghi M. 1984, The ESO Messenger 37, 1 \item Raffi G., Ziebell M. 1986, The ESO Messenger 44, 26 \item Raffi G. et al. 1990, SPIE 1235 \item Tyson N.D., Gal R.R. 1993, AJ, 105, 1206 \item Wallander A. 1993, Observing at a Distance, ed. Emerson D., Clowes R. (World Scientific, Singapore), p. 199 \item Wallander A. 1994, Nuclear Instr. and Methods in Physics Research Vol. A347, 258 \end{list} \end{document}
\section{\label{INTRO} INTRODUCTION} The compounds SrCu$_2$O$_3$ \cite{takano} and (VO)$_2$P$_2$O$_7$ \cite{johnston} consist of arrays of weakly interacting two--chain metal--oxide ladders. In SrCu$_2$O$_3$, CuO$_2$ ladders are coupled by a weak, frustrated, ferromagnetic coupling\cite{rice}, and in (VO)$_2$P$_2$O$_7$, VO$_4$ ladders are well separated in the structure of the material. These materials have a half--filled conduction band, corresponding to one conduction electron per metal--oxide unit and are insulators with short--range antiferromagnetic correlations and a spin gap. Very recently Z.\ Hiroi and M.\ Takano\cite{hit95} succeeded in doping the two--leg ladder compound La$_{1-x}$Sr$_x$CuO$_{2.5}$. They observe a fall in resistivity with doping, possibly to a metallic state, but no sign of superconductivity. The spin--gap found in half--filled LaCuO$_{2.5}$ persists, but decreases with doping. In this publication we will address the half--filled situation only. The undoped materials have been described by a Heisenberg model on two coupled chains \cite{rice,dagotto,barnes1,whiteprl}, which explains a number of experimentally observed magnetic properties of the materials, including the size of the spin gap, and the spectrum of the spin triplet excitations. The Hubbard model on two coupled chains, which at half--filling and in the limit of strong on--site Coulomb repulsion can be mapped onto the Heisenberg model, gives an itinerant electron picture of these systems \cite{nws94} which includes charge as well as spin fluctuations. While the ground state and the low--energy properties of the two--chain Heisenberg and half--filled Hubbard model are becoming well understood, less is known about the finite frequency one-- and two--particle response. The dynamic spin response of the Heisenberg model was studied in Ref.\ 5 and the single particle and spin response was examined for a $t$--$J$ model with two holes in Ref.\ 8 using analytic continuation of Lanczos exact diagonalization calculations on small clusters. The dynamic properties are important because they can be measured with a variety of experimental techniques. For example, the one--particle spectral weight can be probed by photoemission and inverse photoemission, the dynamic charge correlation function by optical response measurements, and the dynamic spin structure factor by inelastic neutron scattering. Here we examine the dynamic one-- and two--particle correlation functions of the two--chain Hubbard model at half--filling using Quantum Monte Carlo simulations analytically continued to real frequencies using a Maximum Entropy technique\cite{maxent}. We study the dynamic response for large, intermediate, and small values of the perpendicular hopping $t_\perp$. This is interesting for two reasons: first, when the electron--electron interaction is turned off ($U=0$) in the two--chain system, there is a metal--insulator transition from a two--band metal to a band insulator due to the separation of the bonding and antibonding bands with increasing $t_\perp$. For the interacting ($U \ne 0$) case, it is generally believed that the two--chain system is always insulating (see, for example, Ref. 11 for a discussion within a weak--coupling RG picture). The QMC simulations do find an insulator for all $t_\perp$, but also reveal a crossover from four--band insulating behavior for small $t_\perp$ to two--band behavior for \mbox{$t_\perp/t$ {\raise3pt\hbox{$>$}\llap{\lower3pt\hbox{$\sim $}}} $2$}. A similar crossover occurs for weak coupling ($U$ {\raise3pt\hbox{$<$}\llap{\lower3pt\hbox{$\sim $}}} $t$) within antiferromagnetic Hartree--Fock (AFHF) theory, as we shall discuss in section \ref{AFHFSEC}. One interesting and important result of this paper is that this four--band to two--band crossover scenario survives in the intermediate to large $U$ regime in the system, despite the fact that the many--body physics for $t_\perp$ {\raise3pt\hbox{$>$}\llap{\lower3pt\hbox{$\sim $}}} $t$ is not even qualitatively reproduced in the AFHF approximation. No such crossover is present as a function of $J_\perp/J$ in the Heisenberg model, because the mapping to the Heisenberg model breaks down when $J_\perp \sim 4 t_\perp^2/U$ is of the order of $t_\perp$. In other words, the effect of the electronic band structure is not present in the Heisenberg model. As we shall see, this crossover can also be described in terms of the crossover from a localized rung picture, in which localized excitations form a Bloch state along the chains for large $t_\perp$, to a picture with longer range antiferromagnetic correlations with many features that can be qualitatively described by AFHF spin--wave theory for small $t_\perp$. Secondly, in the small to intermediate $t_\perp$ regime, this system displays features of the single--particle spectral weight $A({\bf k},\omega)$ seen in recent numerical treatments of both the 1D\cite{preuss1d,newpreuss1d} and 2D\cite{preuss2d,dagotto2d} Hubbard model. We find two features also generic to the 1D and 2D systems: a dispersive band whose width is of the order of the effective exchange interaction $J\sim 4t^2/U$ and an incoherent background of width several times the parallel hopping integral $t$. For the one and two chain systems, but not the 2D system, the dispersive band is well described by renormalized spin--wave--theory. We consider the single band Hubbard model on two coupled chains of length $L$: \begin{eqnarray} H&=&-t\sum_{i,\lambda\sigma} (c^\dagger_{i,\lambda\sigma} c^{\phantom{\dagger}}_{i+1,\lambda\sigma} + h.c.) \nonumber\\ && -t_\perp \sum_{i,\sigma} (c^\dagger_{i,1\sigma} c^{\phantom{\dagger}}_{i,2\sigma} + h.c.) +U\sum_{i\lambda} n_{i,\lambda\uparrow} n_{i,\lambda\downarrow}. \label{hamiltonian} \end{eqnarray} Here $c^\dagger_{i,\lambda\sigma}$ and $c^{\phantom{\dagger}}_{i,\lambda\sigma}$ create and annihilate electrons on rung $i$ and chain $\lambda$ with spin $\sigma$, respectively, the hopping integral parallel to the chains is $t$, the hopping between the chains $t_\perp$, and $U$ is an on-site Coulomb interaction. We use periodic boundary conditions parallel to and open boundary conditions perpendicular to the chains. The non--interacting, $U=0$, Hamiltonian can be diagonalized by writing the hopping term in terms of bonding and antibonding states on a rung and Fourier--transforming parallel to the chains. The energy is then given by \begin{equation} \varepsilon_{\bf k} = -(2t \cos k + t_\perp \cos k_\perp) \label{eqepsilon} \end{equation} with ${\bf k}=(k,k_\perp)$, where $k_\perp=0$ and $k_\perp=\pi$ corresponds to the energy of the bonding and antibonding band, respectively, and $k$ is the momentum along the chains. Both bands will be occupied when $t_\perp < t_{\perp c}$, whereas when $t_\perp > t_{\perp c}$, only the bonding band will be occupied. Here $t_{\perp c}/t \equiv 1 - \cos \pi \langle n \rangle$ and $\langle n \rangle \equiv \langle \sum_\sigma c^\dagger_{i,\lambda,\sigma} c_{i,\lambda,\sigma}\rangle$. For half--filling $t_{\perp c}/t=2$, and the system is a two--band metal for $t_\perp/t < 2$, and a band insulator with a completely occupied bonding band for $t_\perp/t > 2$. \section{\label{SINGLE} SINGLE--PARTICLE SPECTRAL WEIGHT} \subsection{ Quantum Monte Carlo Results} In order to calculate the dynamical properties of the above model we employ the grand--canonical Quantum Monte Carlo (QMC) algorithm to determine the expectation values of the correlation functions in imaginary time. At half--filling, the simulations are not limited by the fermion sign problem and accurate results can thus be achieved at low temperatures and large system sizes. We analytically continue the imaginary--time data to real frequencies using a Maximum Entropy method\cite{maxent}. For example, the spectral weight function given by \begin{eqnarray} A({\bf k},\omega) &=& {1\over Z} \sum_{l,l^\prime} e^{-\beta E_l}(1+e^{-\beta\omega}) \vert \langle l\vert c_{{\bf k},\uparrow}\vert l^\prime \rangle \vert^2 \nonumber\\ && \phantom{{1\over Z}\sum_{l,l^\prime}} \cdot\delta(\omega-(E_{l^\prime}-E_l)), \label{Akomegadef} \end{eqnarray} where $Z$ is the partition function, $\vert l>$ and $\vert l^\prime>$ are the exact many-body eigenstates with the corresponding energies $E_l$ and $E_{l^\prime}$ and $c_{{\bf k},\sigma}=\sum_{i,\lambda} c_{i,\lambda\sigma} e^{i{\bf k}\cdot{\bf R}_{i,\lambda}}/\sqrt{2L}$, can be calculated by inverting the spectral theorem \begin{equation} G({\bf k},\tau)\equiv \langle c^{\phantom{\dagger}}_{{\bf k},\uparrow}(\tau) c^\dagger_{{\bf k},\uparrow}(0) \rangle = \int_{-\infty}^{\infty} {e^{-\tau\omega}\over 1+e^{-\beta\omega}} A({\bf k},\omega) d\omega \label{eqGtaudef} \end{equation} with the Maximum Entropy algorithm using the raw QMC data $G({\bf k},\tau)$. To achieve high resolution, it is important to use a likelihood function which takes the error--covariance matrix of the QMC data and its statistical inaccuracy consistently into account\cite{preussmaxent}. The results presented here are based on QMC data with good statistics, i.e. averages over $10^5$ updates of all the Hubbard--Stratonovich variables result in $G({\bf k}, \tau)$'s with absolute errors less than or of the order of $5 \times 10^{-4}$. Correlations of the data in imaginary time were taken into account by making use of the covariance matrix in the Maximum Entropy procedure \cite{preussmaxent}. As suggested in previous work by White\cite{whi91}, various moments of the spectral weight were also incorporated in extracting $A({\bf k},\omega)$. We calculate the spectral function $A({\bf k},\omega)$ at half--filling for a $2\times 16$-lattice with $U/t=8$ for different values of $t_\perp$ at an inverse temperature of $\beta=10/t$. Our results are given in Fig.\ \ref{figqmc2.0} for $t_\perp/t=2.0$, in Fig.\ \ref{figqmc1.0} for $t_\perp/t=1.0$, and in Fig.\ \ref{figqmc0.5} for $t_\perp/t=0.5$. Note that $t_\perp/t=1.0$ corresponds to the isotropic case, $J_\perp\approx J$, relevant to the SrCu$_2$O$_3$, (VO)$_2$P$_2$O$_7$, and LaCuO$_{2.5}$ compounds. Parts (a) and (b) in Fig.\ \ref{figqmc2.0}--\ref{figqmc0.5} are three-dimensional plots of $A({\bf k},\omega)$ versus $\omega$ for $k$--values in the 1D Brillouin zone, whereas part (c) in all three figures summarize these results in the usual ``band structure'' $\omega$ versus $k$ plot. The shaded areas represent regions with significant spectral weight, with the darker shading representing more weight. Due to particle--hole symmetry, $A({\bf k}=(k,\pi),\omega)=A({\bf k}=(\pi-k,0),-\omega)$ where ${\bf k}=(k,k_\perp)$. In other words, one reflects $k$ about $k=\pi/2$ and $\omega$ about 0 to get $A({\bf k},\omega)$ for $k_\perp=\pi$ from $A({\bf k},\omega)$ for $k_\perp=0$. This symmetry can be seen by comparing parts (a) and (b) of Fig.\ \ref{figqmc2.0}--\ref{figqmc0.5}. Since this symmetry is not enforced by the Maximum Entropy procedure, it is only present approximately, and the extent to which the reflected spectra at $k_\perp=0$ do not match those at $k_\perp=\pi$ gives an indication of the accuracy of the analytic continuation procedure. In the density plots in parts (c), we show only the $k_\perp=0$ components. For $t_\perp/t=2.0$, there is a heavily weighted coherent band of width $\sim 3t$ in the photoemission ($\omega < 0$) part of the spectrum. In the inverse photoemission spectrum ($\omega > 0$) there is very little total spectral weight. There are therefore two bands with significant spectral weight, one in the photoemission spectrum for $k_\perp=0$ and one in the inverse photoemission spectrum for $k_\perp=\pi$. Each band is about $2t$ away from the Fermi surface, leading to a gap of approximately $4t$. Therefore, for large $t_\perp$, the system is a two-band insulator. In contrast, for $t_\perp/t=1.0$ and $t_\perp/t=0.5$, $A({\bf k},\omega)$ has substantial spectral weight in four bands, one in the photoemission spectrum ($\omega<0$) for $k_\perp=0$ and $k_\perp =\pi$ and one in the inverse photoemission spectrum ($\omega >0$) for both values of $k_\perp$. In this regime, the system is thus a four--band insulator. For $t_\perp/t=0.5$ (Fig.\ \ref{figqmc0.5}) the spectral weight is present for both values of $k_\perp$ concentrated between $k=0$ and $k=\pi/2$ in the photoemission part and between $k=\pi/2$ and $k=\pi$ in the inverse photoemission part of the spectrum. For the isotropic case ($t_\perp/t=1.0$, Fig.\ \ref{figqmc1.0}), there is some shift of spectral weight to the photoemission part for $k_\perp=0$ and to the inverse photoemission part for $k_\perp=\pi$. The maxima of the photoemission band in the $k_\perp=0$ part occurs at $k^*=\pi$ for $t_\perp/t=2.0$, at $k^*\approx 0.7\pi$ for $t_\perp/t=1.0$, and at $k^*=\pi/2$ for $t_\perp/t=0.5$. Therefore we would expect to see a maxima at $k^*\approx 0.7\pi$ in a photoemission experiment on the isotropic ladder compounds, an experiment which has, to our knowledge, not yet been done. It is also important to note that the QMC $A({\bf k},\omega)$ results for $t_\perp/t=0.5$ as well as for the isotropic case, $t_\perp/t=1.0$, display two general features which have recently been observed in both the 1D\cite{newpreuss1d} and 2D\cite{preuss2d} cases using the improved Maximum Entropy techniques described above\cite{preussmaxent}. One feature is that $A({\bf k},\omega)$ contains a rather dispersionless ``incoherent background'' extending over several $t$ ($\sim 6t$ for $U/t=8$ in 2D) in both the electronically occupied ($\omega < 0$) and unoccupied ($\omega > 0$) parts of the spectrum. The crucial structure, not resolved in previous 1D and 2D QMC simulations, is a dispersive structure at low energies with a small width of the order of the exchange coupling $J=4t^2/U$. It is this latter coherent ``band'' which defines the gap $\Delta$ and which, for larger $U$ ($U/t\sim 10$), is well--separated from the higher--energy background. As in the 1D and 2D cases, the splitting in the low--energy band and the higher--energy background is especially pronounced near $k=0$ and $k=\pi$ due to a relative weight shift from negative to positive energies as $k$ moves through $k^*$. \subsection{\label{AFHFSEC}Spin--Wave Theory} At half--filling, the two chain Hubbard system at $T=0$ is a spin liquid with a spin gap and a finite spin--spin correlation length. The system is therefore less ordered than either the 1D system, which is at the critical point and has power law decay of the spin--spin correlation function and no spin gap, or the 2D system, which has long--range spin order and gapless spin--wave--like excitations. However, at large $U/t$ the scale of the spin gap for the two chain system is set by $J_\perp \sim 4t_\perp^2/U$, and the spin--spin correlation length $\xi$ grows longer as the spin gap gets smaller. The range of the spin ordering can therefore be tuned by varying $t_\perp/t$. This is illustrated by Fig.\ \ref{figDMRG}, in which we plot the spin--spin correlation function, $S(r)=(-1)^r \langle M_{0,\lambda}^z M_{r,\lambda}^z \rangle$ on a semilog scale for $t_\perp/t=2.0$, $t_\perp/t=1.0$ and $t_\perp/t=0.5$. Here $M_{r,\lambda}^z=(n_{r,\lambda,\uparrow} - n_{r,\lambda,\downarrow})$ is twice the $z$ component of the on--site spin. These calculations were done using the Density Matrix Renormalization Group method\cite{nws94} (DMRG) on a system with open boundaries at zero temperature. For $t_\perp/t=0.5$, the data was averaged over a number of different pairs of points for a given $r$ in order to reduce oscillations due to the open boundaries. For distances greater than a few lattice spacings, the correlation functions are very well fit by a pure exponential form $A\exp(-r/\xi)$, from which the spin--spin correlation length $\xi$ can be determined. For $t_\perp/t=2.0$, $\xi=0.83$, for $t_\perp/t=1.0$, $\xi=4.3$, and for $t_\perp/t=0.5$, $\xi=9.5$ lattice spacings. Therefore, the spins are almost completely uncorrelated along the chains for $t_\perp/t=2.0$, but for $t_\perp/t=1.0$ and for $t_\perp/t=0.5$ are correlated on length scales of the order of the size of the $2\times 16$ lattice studied here, although the long--range behavior is that of a gapped spin liquid in all three cases. It is therefore reasonable to calculate the single particle spectral weight $A({\bf k},\omega)$ in two simple ways: first, when the spin gap is small and $\xi \approx L$, and we consider wavelengths smaller than $\xi$ and frequencies larger than the spin gap, it is useful to explore the consequences of a simple SDW approximation based on an ordered state, i.e. AFHF theory. Secondly, one can solve for the exact eigenstates on a rung and then consider states composed of a product of exact rung singlets only and a delocalized one--hole state in a background of rung singlets. (We call this the Local Rung Approximation, LRA.) The LRA starts with a state with uncorrelated rungs and then treats the interactions between rungs perturbatively, and is thus a good starting point when $\xi$ is small. As we shall see, the LRA gives a single particle spectral weight distribution that agrees well with QMC in the large $t_\perp$ regime, while AFHF qualitatively describes most features of the spectral weight distribution for small $t_\perp$. We first discuss the AFHF calculation in more detail. The Hartree-Fock one--particle Hamiltonian can be written as \begin{equation} H_{HF}={\sum_{{\bf k},\sigma}}^\prime E_{\bf k} (b^{c\dagger}_{{\bf k},\sigma} b^{c\phantom{\dagger}}_{{\bf k},\sigma} - b^{v\dagger}_{{\bf k},\sigma} b^{v\phantom{\dagger}}_{{\bf k},\sigma}), \end{equation} where ${\sum_{{\bf k},\sigma}}^\prime$ means that the sum runs only over the magnetic zone ($-\pi/2 < k\le \pi/2$, $k_\perp=0,\pi$). The operators $b^c_{{\bf k},\sigma}$ and $b^v_{{\bf k},\sigma}$ are given by the following transformation\cite{swz89}: \begin{eqnarray} b^c_{{\bf k},\sigma} & = & u_{\bf k} c_{{\bf k},\sigma} \mp v_{\bf k} c_{{\bf k}+{\bf Q},\sigma},\\ b^v_{{\bf k},\sigma} & = & v_{\bf k} c_{{\bf k},\sigma} \pm u_{\bf k} c_{{\bf k}+{\bf Q},\sigma}, \end{eqnarray} where ${\bf Q}=(\pi,\pi)$, the upper (lower) sign corresponds to $\sigma=\uparrow$ ($\sigma=\downarrow$), and \begin{eqnarray} u_{\bf k} & = & \sqrt{{1\over 2}(1+\varepsilon_{\bf k}/E_{\bf k})},\\ v_{\bf k} & = & \sqrt{{1\over 2}(1-\varepsilon_{\bf k}/E_{\bf k})},\\ E_{\bf k} & = & \sqrt{\Delta^2+{\varepsilon_{\bf k}}^2}. \label{eqAFHFdisp} \end{eqnarray} Here $\varepsilon_{\bf k}$ is the free particle energy ($U=0$) defined in Eq.\ (\ref{eqepsilon}). The spin--density--wave gap, $\Delta$, is self--consistently determined by the equation \begin{equation} {1\over U} = {1\over 2L}{\sum_{\bf k}}^\prime {1\over E_{\bf k}}. \label{eqAFHFgap} \end{equation} The band structure of the AFHF Hamiltonian is given by $\pm E_{\bf k}=\pm\sqrt{\Delta^2+{\varepsilon_{\bf k}}^2}$ in the magnetic Brillouin zone. The SDW gap $\Delta$, calculated by solving Eq.\ (\ref{eqAFHFgap}) self--consistently, is shown as a function of $t_\perp/t$ for $U/t=2$, 4, and 8 in Fig.\ \ref{figAFHFgap}. The transition from a two--band metal to a band insulator in the $U=0$ system occurs at $t_\perp/t=t_{\perp c} = 2.0$. As $U$ is turned on for $t < t_{\perp c}$ in the AFHF, a gap develops in both the bonding and antibonding bands, leading to a four--band insulator. If $t_\perp$ is then increased, $\Delta$ will go to zero at approximately the noninteracting ($U=0$) $t_{\perp c}/t=2.0$, as seen for $U/t=2$ in Fig.\ \ref{figAFHFgap}, leading to a transition from a four--band to a two--band insulator. For large $U$ and small $t_\perp$, $\Delta \approx U/2$, as seen for $U/t=8$ in Fig.\ \ref{figAFHFgap}, so the Coulomb splitting of the bands will dominate over $U=0$ band structure effects. In this case, the gap will go to zero, i.e. the system will undergo a transition to a band insulator only when the band structure splitting $2 t_\perp$, is of the order of the Coulomb splitting, so that $t_{\perp c} \approx \Delta \approx U/2$. The spectral weight within the AFHF approximation is \begin{equation} A({\bf k},\omega) = u_{\bf k}^2 \delta (\omega - E_{\bf k}) + v_{\bf k}^2 \delta (\omega + E_{\bf k}). \end{equation} We show the spectral weight $A({\bf k},\omega)$ from AFHF for the same parameters as in Fig.\ \ref{figqmc2.0} and Fig.\ \ref{figqmc0.5} in Fig.\ \ref{figAFHF}(a) for $t_\perp/t=2.0$ and Fig.\ \ref{figAFHF}(b) for $t_\perp/t=0.5$. The band splitting $2 \Delta \approx U$ in both cases, as can also be seen in Fig.\ \ref{figAFHFgap}. Since both the bonding and antibonding bands are split by this gap, there are two peaks in $A({\bf k},\omega)$ as a function of $\omega$ for each ${\bf k}$, leading to a four--band insulator for both values of $t_\perp/t$. Since the band splitting is set by $U$ rather than $2 t_\perp$, the spectral weight is almost evenly distributed between two coherent bands for both $t_\perp/t=2.0$ and $t_\perp/t=0.5$. Therefore, for $t_\perp/t=2.0$, as can be seen by comparing the QMC results in Fig.\ \ref{figqmc2.0} with Fig.\ \ref{figAFHF}(a), the AFHF spectral weight distribution for $t_\perp/t=2.0$ is completely wrong, with too much weight in the upper, $\omega>0$, band for $k_\perp=0$ and the $\omega < 0$ band for $k_\perp=\pi$. While the average positions of the AFHF bands, shown as thick solid lines in Fig.\ \ref{figqmc2.0}(c), are approximately the same as those of the QMC bands, the band widths are too small by approximately a factor of two. For $t_\perp/t=2.0$ and $U/t=8$, then, the AFHF calculation results in a four--band insulator in which there is long--range antiferromagnetic order, while the QMC results show that the system is a two--band insulator, with only very short range antiferromagnetic correlations. While the AFHF picture does produce a two--band insulator for $t_\perp/t$ {\raise3pt\hbox{$>$}\llap{\lower3pt\hbox{$\sim $}}} $4.5$, the physical picture of the transition to this phase is quite different (see section C). For $t_\perp/t=0.5$, as seen in Fig.\ \ref{figqmc0.5} and Fig.\ \ref{figAFHF}(b), AFHF gives two dispersive bands in the $k_\perp=0$ branch of width of order $J \sim 4 t^2/U$, similar to the coherent bands seen in the QMC data. However, in the AFHF, the single particle gap is somewhat too large, the weight distribution extends too far towards $k=\pi$ and too far towards $k=0$ for the photoemission and inverse photoemission parts of the spectrum, respectively, and the broad incoherent background is not present. We have also carried out mean--field slave boson calculations\cite{slaveboson} in order to try to improve on the AFHF picture. These calculations give a gap about 15\% smaller than AFHF, and a bandwidth about 3\% smaller, but do not qualitatively improve on the AFHF calculations. The same holds for $t_\perp/t=1.0$, where the AFHF describes approximately the coherent part of the spectrum and also produces a too large single particle gap and a slightly different spectral weight distribution. Therefore, for $t_\perp/t=0.5$ and for $t_\perp/t=1.0$, AFHF would give a reasonable description of the dispersion and general spectral weight distribution of the coherent spin--wave bands, if the gap were phenomenologically adjusted to a smaller value, but fails to even qualitatively describe the spectral weight distribution for $t_\perp/t=2.0$. \subsection{\label{RUNGSINGLET} Local Rung Approximation} For very large $t_\perp$, a better starting point is the limit of weak interaction between the rungs. In this limit, we split the Hamiltonian of Eq.\ (\ref{hamiltonian}) into $H=H_0 + H_I$ with \begin{eqnarray} H_0&=&-t_\perp \sum_{i,\sigma} (c^\dagger_{i,1\sigma} c^{\phantom{\dagger}}_{i,2\sigma} + h.c.) + U\sum_{i\lambda} n_{i,\lambda\uparrow} n_{i,\lambda\downarrow}, \nonumber\\ H_I&=&-t\sum_{i,\lambda\sigma} (c^\dagger_{i,\lambda\sigma} c^{\phantom{\dagger}}_{i+1,\lambda\sigma} + h.c.), \end{eqnarray} where $H_0$ denotes the non--interacting rung limit. The ground state of $H_0$ is a product of individual rung eigenstates. Accordingly, we first diagonalize the Hubbard Hamiltonian on the two sites making up the rung in order to find the rung eigenstates. A schematic diagram of the exact eigenstates for a single rung is shown in Fig.\ \ref{figrunglevels}. The ground state for the half--filled rung $i$ (two particles per rung) is a spin singlet state with $k_\perp=0$ and energy $E_0/L = E_0^{\text{rung}} = \left( U- \sqrt{U^2+ 16 t_\perp^2} \right)/2$ denoted by $\vert S_i\rangle$. For large values of $U$, the energy of this two-site state is given by the exchange coupling $-J_\perp \sim -4t_\perp^2/U$. Note that $\vert S_i\rangle$ contains terms which have double occupied sites as well as the usual singlet construction $(\vert\uparrow,\downarrow\rangle -\vert\downarrow,\uparrow\rangle)/\sqrt 2$. We can then form an approximation to the ground state for the half--filled system by taking a state which is just a product of the rung states, \begin{equation} |\psi_0\rangle = | S_1 \rangle | S_2 \rangle ... | S_L \rangle . \label{eqsingstate} \end{equation} For $t_\perp/t=2.0$, the binding energy of a singlet formed on a rung is about four times lower than the energy of a singlet between neighboring sites along a chain, and therefore we expect $|\psi_0\rangle$ to be a good approximation to the exact ground state. In order to obtain the $\omega<0$ spectral weight for the half--filled system, we need to calculate the matrix element of $c_{{\bf k}, \sigma}$ given in Eq.\ (\ref{Akomegadef}). In other words, we need matrix elements between a half--filled state and a state with one particle removed. We form the approximate one hole state by replacing one bond singlet state at rung $\ell$, $|S_\ell\rangle$, with the lowest energy one--particle state, the one with bonding ($k_\perp=0$) symmetry, $|B_{\ell\uparrow}\rangle$ (We remove a spin down electron for definiteness.) and define \begin{equation} |\ell\rangle = | S_1 \rangle | S_2 \rangle ... |B_{\ell\uparrow}\rangle ...| S_L \rangle. \label{eqlocalsing} \end{equation} We delocalize the single--particle state with a plane wave ansatz by constructing the state\cite{barnes1,bos93} \begin{equation} |\psi_1(k)\rangle = L^{-1/2}\sum_{\ell=1}^L e^{ik\ell} |\ell\rangle. \label{eqdelocalsing} \end{equation} Trivially, this state is an exact eigenstate of $H_0$. We will take this state as the ground state for the system with one hole with momentum ${\bf k} = (k,0)$. The spectral weight $A({\bf k},\omega)$ at zero temperature ($\beta\rightarrow\infty$) is then approximated by using only the two states $|\psi_0\rangle$ and $|\psi_1(k)\rangle$ in Eq.\ (\ref{Akomegadef}). For $k_\perp=0$, the energy dispersion $\omega(k)$ of $A({\bf k},\omega)$ for $\omega < 0$ is given by \begin{eqnarray} \omega(k) &=&\langle\psi_0\vert H \vert\psi_0\rangle - \langle \psi_1(k)\vert H \vert\psi_1(k) \rangle - \mu \nonumber\\ &=&-\frac{1}{2} \sqrt{U^2+ 16 t_\perp^2} + t_\perp -tA \cos k, \label{eqLRAwk} \end{eqnarray} with $A=(1+E_2/2t_\perp)^2/(1 + E_2^2/4t_\perp^2)$, $E_2=( U + \sqrt{U^2+ 16 t_\perp^2})/2$, and the corresponding spectral weight by $\vert\langle\psi_1(k)\vert c_{k,\downarrow} \vert\psi_0 \rangle\vert^2$. The dispersion given by Eq.\ (\ref{eqLRAwk}) for $U/t=8$ and $t_\perp/t=2.0$ is plotted in Fig.\ \ref{figqmc2.0}(c) and the spectral weight for $k_\perp=0$ is shown in Fig.\ \ref{figLRA}. One can see that the position of the peak of the $\omega < 0$ LRA band, denoted LRA1 in Fig.\ \ref{figqmc2.0}(c), lays almost exactly on the QMC data. However, the position and the dispersion of the $\omega > 0$ band, which were obtained in the same way as for $\omega<0$, are not exactly matched by the LRA1 calculation. While this band is not very important in the sense that it contains very little spectral weight, we can understand how to improve the LRA1 calculation by considering the states of a four site cluster. In order to generate the inverse photoemission spectrum, we need to calculate matrix elements between the half--filled state and a state with one {\it additional} particle. {}From Fig.\ \ref{figrunglevels}, we see that a rung state with three particles and $k_\perp=0$ has approximately twice the excitation energy of the $k_\perp=0$ one--particle state which is relevant for the photoemission part of the spectrum. (Recall that due to the particle--hole symmetry at half--filling, this state will map to one in the inverse photoemission part of spectrum for $k=\pi$; one can see this symmetry in Fig.\ \ref{figrunglevels}.) Since the relevant three--particle single rung state is high in energy, configurations involving intrachain effects might also be important. We include the effect of such configurations by replacing two rung singlet states by the lowest energy state of five particles on four sites in the bonding channel. We can then form a delocalized plane wave state from this state as in Eq.\ (\ref{eqdelocalsing}). The results of this calculation are labeled as LRA2 in Fig.\ \ref{figqmc2.0}(c). The location and width of the $k_\perp=0$ inverse photoemission band are more closely fit than by the LRA1 calculation. In Fig.\ \ref{figLRA} the spectral weight distribution of the LRA1 calculation for the photoemission part of the spectrum ($\omega<0$) and the distribution of the LRA2 calculation for the inverse photoemission part ($\omega>0$) is plotted. The result is in good accordance with the QMC data in Fig.\ \ref{figqmc2.0}(a). \section{\label{SPINCHARGE} SPIN AND CHARGE DYNAMIC CORRELATION FUNCTIONS} In order to determine the nature of the low lying excitations, we also consider the spin and charge susceptibilities $\chi_{s,c}({\bf q},\omega)$ which, in a Lehmann representation, are defined as \begin{eqnarray} \chi_{s,c}({\bf q},\omega)&=& {i\over Z}\sum_{l,l^\prime}e^{-\beta E_l} (1-e^{-\beta\omega}) \vert\langle l\vert O_{s,c}({\bf q})\vert l^\prime\rangle\vert^2 \nonumber\\ &&\phantom{{i\pi\over Z}\sum_{l,l^\prime}} \cdot\delta(\omega-(E_{l^\prime}-E_l)), \label{eqtwocorrdef} \end{eqnarray} with $O_{s}({\bf q})=\sum_{\bf p} (c^\dagger_{{\bf p}+{\bf q},\uparrow}c^{\phantom{\dagger}}_{{\bf p},\uparrow} -c^\dagger_{{\bf p}+{\bf q},\downarrow} c^{\phantom{\dagger}}_{{\bf p},\downarrow})$ and $O_{c}({\bf q})=\sum_{{\bf p},\sigma} c^\dagger_{{\bf p} +{\bf q},\sigma}c^{\phantom{\dagger}}_{{\bf p},\sigma}$. We calculate the two--particle dynamic response from the two--particle imaginary time Green's function, as in Eq.\ (\ref{eqGtaudef}) and use the Maximum Entropy method described above for the analytical continuation of $\chi_{s,c}({\bf q},\omega)$ to real frequencies. In Fig. \ref{figspin} we show the dynamical spin--spin correlation function $\chi_{s,c}({\bf q},\omega)$ for $t_\perp/t=2.0$ and $t_\perp/t=1.0$. For $t_\perp/t=2.0$, the $q_\perp=0$ component [Fig.\ \ref{figspin}(a)] has a broad, lightly weighted structure centered approximately at $\omega=3t$, with the spectral weight vanishing for small $q$. For $q_\perp=\pi$ [Fig.\ \ref{figspin}(b)], there is a coherent, dispersive band with a width set by $J \approx 4 t^2/U = 0.5t$, a minimum at $q=\pi$ and a maximum at $q=0$. The minimum spin gap is approximately $0.8t$, which agrees well with the value we obtain using Projector Quantum Monte Carlo, indicated by the dashed line on the plot. The spectral weight is most heavily concentrated around $q=\pi$. The dynamic spin response at half--filling measures the response of the system to spin triplet excitations. Since here $t_\perp$ is large, we consider the effect of triplet excitations on product rung states of the type considered in Eq.\ (\ref{eqsingstate}), in the LRA calculation. A triplet excitation will change the total spin of the state from $S=0$ to $S=1$. Referring to Fig.\ \ref{figrunglevels}, the only triplet excited state on a single rung has momentum $\pi$ (corresponding to odd parity under chain interchange) leading to a change in momentum $q_\perp=\pi$ for the triplet excitation. The size of this triplet excitation $\Delta E^{\text{rung}}_{\text{s}}$ is marked on Fig.\ \ref{figspin}(b) by the solid line. This local triplet excited state can be moved to a near neighbor rung by a process which is second order in $H_I$. The local triplet excited states can then be delocalized into a Bloch wave as in Eq.\ (\ref{eqlocalsing}) and Eq.\ (\ref{eqdelocalsing}). For large $U$, one then obtains a dispersion relation of the form $\Delta E^{\text{rung}}_{\text{s}} + J \cos q$, where $J\sim 4t^2/U$, a form similar to that obtained in Ref.\ 5 for the two chain Heisenberg model. The coherent band in Fig.\ \ref{figspin}(b) does have a minimum at $q=\pi$, and has approximately this form. Within the LRA picture, it is also possible to excite a $q_\perp=0$ (even under chain interchange) triplet excitation via a ``two--magnon'' process, as discussed in Ref.\ 5. This corresponds to making two local triplet excitations on rungs in the non--interacting rung picture, and would lead to an excitation energy whose scale, in the Heisenberg limit, is set by $\omega \sim 2J_\perp$ rather than $\omega \sim J_\perp$. The overall position of the lightly weighted band seen for $q_\perp=0$ in Fig.\ \ref{figspin}(a) is consistent with this two--magnon band, although the band is too lightly weighted and our resolution too low to extract a dispersion relation. For the physical relevant isotropic case ($t_\perp/t=1.0$), shown in Fig.\ \ref{figspin}(c) and (d), the spin response looks quite different. (Because the $t_\perp/t=0.5$ results are qualitatively similar to the $t_\perp/t=1.0$ results, we show only the isotropic ($t_\perp/t=1.0$) case here.) In Fig.\ \ref{figspin}(c) and (d), there are dispersive bands in both the $q_\perp=0$ and $q_\perp=\pi$ branches, with the position of the peak going to a finite minimum at ${\bf q}=(0,\pi)$ and ${\bf q}=(\pi,0)$, and appears to vanish as ${\bf q} \rightarrow (\pi,\pi)$. However, as shown in Fig.\ \ref{figDMRG}, the correlation length is 4.3 lattice spacings in this regime, and there should be a spin gap. Using DMRG calculations on lattices of $2 \times 8$ to $2 \times 32$ sites\cite{nws94}, we estimate the spin gap to be of order $0.12t$. Due to finite size effects and finite resolution in the analytic continuation, this gap is too small to resolve in Fig.\ \ref{figspin}(d). Near ${\bf q}=(0,0)$ [Fig.\ \ref{figspin}(c)], the dispersion of the peak is hard to discern because there is very little spectral weight. This lack of spectral weight at ${\bf q}=(0,0)$ is present in all the dynamic charge and spin correlations and is due to a selection rule that comes about because $O_{s,c}({\bf q=0})$ in Eq.\ (\ref{eqtwocorrdef}) commutes with the Hamiltonian, leading to a vanishing of the matrix element of $O_{s,c}({\bf q})$ as ${\bf q} \rightarrow 0$. Therefore, the system has relatively long--range spin order in this regime and, to within the resolution of our calculations, shows the characteristics of an ordered state with gapless excitations. Thus it is interesting to compare the QMC spectra to the results of spin--wave theory calculations. In order to extract the low--lying spin excitations using spin--wave theory it is necessary to consider the RPA transverse spin susceptibility $\chi^{+-}_{\rm RPA}({\bf k},{\bf k}^\prime,\omega)$. It is obtained by applying the random--phase approximation to the response function $\chi^{+-}_0({\bf k},{\bf k}^\prime,\omega)$, which is calculated directly in real time using the AFHF ground state $\vert\Omega\rangle$, calculated in section \ref{AFHFSEC}: \begin{equation} \chi^{+-}_0({\bf q},{\bf q}^\prime,t)= {i\over 4L} \langle\Omega\vert T S^+_{\bf q}(t)S^-_{-{\bf q}^\prime} (0)\vert\Omega\rangle, \end{equation} with $S^\pm=S_x\pm iS_y$. Due to the broken spin rotational symmetry of $\vert\Omega\rangle$, $\chi^{+-}_{\rm RPA}({\bf q},{\bf q}^\prime,\omega)$ contains a gapless mode, as predicted by the Goldstone theorem. Following the procedure of Schrieffer et al.\cite{swz89}, one obtains: \begin{eqnarray} \lefteqn{ \chi^{+-}_{\rm RPA}({\bf q},{\bf q}^\prime,\omega)=}\ \ \ \ \ \nonumber\\ &&\sum_{\bf p} \chi_0^{+-}({\bf q},{\bf p},\omega) [1-U\chi_0^{+-}({\bf p},{\bf q}^\prime,\omega)]^{-1}, \end{eqnarray} where $[1-U\chi_0^{+-}({\bf p},{\bf q}^\prime,\omega)]^{-1}$ is a matrix inverse of a $2\times 2$ matrix in ${\bf q}$-space and \begin{eqnarray} \chi^{+-}_0({\bf q},{\bf q}^\prime,\omega)&=& \delta({\bf q}-{\bf q}^\prime)\chi_0^{+-}({\bf q},\omega)\nonumber\\ &&+ \delta({\bf q}-{\bf q}^\prime+{\bf Q})\chi_Q^{+-}({\bf q},\omega), \end{eqnarray} with $\chi^{+-}_0({\bf q},\omega)$ and $\chi^{+-}_Q({\bf q},\omega)$ given by the usual ``bubble'' diagrams and printed in detail in Ref.\ 16 [up to a misprint of a factor of 2 in $\chi^{+-}_Q({\bf q},\omega)$]. The dispersion of $\chi^{+-}_{\rm RPA}({\bf q},{\bf q},\omega)$ is indicated as a solid line in Fig.\ \ref{figspin}(c) and \ref{figspin}(d). The RPA spin--wave dispersion goes to zero at ${\bf q}=(0,0)$ and ${\bf q}=(\pi,\pi)$, and goes to a finite minimum at ${\bf q}=(0,\pi)$ and ${\bf q}=(\pi,0)$, consistent with the QMC data. The spin--wave velocity, the width of the bands, and the gaps at ${\bf q}=(0,\pi)$ and ${\bf q}=(\pi,0)$ are also in reasonable agreement with the QMC data. The QMC results for the dynamic charge correlation function are shown in Fig.\ \ref{figcharge} for $t_\perp/t=2.0$ and $t_\perp/t=1.0$. For $t_\perp/t=2.0$, almost all of the spectral weight is in the $q_\perp=\pi$ component, so we do not show the $q_\perp=0$ component. There exist two important features: one is a heavily weighted, relatively flat band at $\omega \sim 10t$ with heaviest weight near $q=0$. This band becomes somewhat incoherent as $q$ increases. The second is a flat, dispersive, less heavily weighted band with a minimum of order $5t$ near $q=\pi$. The spectral weight in this band extends from about $q=\pi/4$ to $q=\pi$, and the size of the charge gap is set by this lightly weighted lower band. In the large $t_\perp$ limit, one can understand the structure of the charge response from the LRA picture described in section \ref{RUNGSINGLET}. In the single rung picture, a charge excitation will occur through a transition to an excited state conserving the number of particles on the rung and the total spin; in other words, the important transition will be in the middle column of Fig.\ \ref{figrunglevels}, from the low-lying $S=0$ state to the higher $S=0$ states. There are two possible charge excited states on the rung, one with momentum $k_\perp=0$ (even parity) and one with momentum $k_\perp=\pi$ (odd parity). In a single rung picture, an optical transition from the $k_\perp=0$ ground state to a $k_\perp=0$ excited state is forbidden because the $q_\perp=0$ density operator $O_c(0)$ commutes with the Hamiltonian. This selection rule forbidding a $q_\perp=0$ optical transition remains present when the rung charge excited states are constructed as in Eq.\ (\ref{eqlocalsing}) and delocalized in a state like that in Eq.\ (\ref{eqdelocalsing}). This is why there is almost no spectral weight in the $q_\perp=0$ portion of the charge response for $t_\perp/t=2.0$. Of course, the system is not exactly in a LRA state, so there will be some higher order processes that will introduce a very small amount of spectral weight into the $q_\perp=0$ branch. The energy of the $q_\perp=\pi$ single rung transition, indicated by a solid line on the Fig.\ \ref{figcharge}(a), gives an excitation energy that agrees well with energy of the heavily weighted region at $q=0$. In order to qualitatively understand the origin of the dispersion of the charge response, one can consider the possible particle--hole excitations within the one--particle band structure given by $A({\bf k},\omega)$ in Fig.\ \ref{figqmc2.0}. There will be significant amplitude in the two--particle charge response when there is significant amplitude for a transition at a particular ${\bf q}=(q,q_\perp)$ and $\omega$ for a particle--hole excitation built up from the one--particle spectral weight. For example, to understand the heavily weighted amplitude at ${\bf q} = (0,\pi)$, one has to integrate over all transitions from the photoemission band in Fig.\ \ref{figqmc2.0}(a) to the inverse photoemission band in Fig.\ \ref{figqmc2.0}(b) which transfer this momentum. Since the single--particle bands are sharp and parallel, one should obtain a single well--defined peak for ${\bf q} = (0,\pi)$, which we see in Fig.\ \ref{figcharge}(a) by considering excitations between the $\omega<0$, $k_\perp=0$ band, and the $\omega>0$, $k_\perp=\pi$ band. In addition, the transition at ${\bf q} = (0,\pi)$ has odd parity between the chains, and is thus allowed by the selection rules for the density operator as $q\rightarrow 0$. As the parallel component $q$ is increased, one can see that there will be a continuum of excitation energies which gets wider as $q$ increases. The minimum excitation energy (position of the lower band) from this construction is shown in Fig.\ \ref{figcharge}(a) by the line with solid dots. However, the excitation energy obtained is consistently smaller than the energy of the lowest heavily weighted band from QMC. Using the single particle bands to construct the two--particle excitations is equivalent to calculating the charge response using the lowest order ``bubble'' diagram, but with exact single--particle propagators, neglecting all particle--hole interactions. The particle--hole interactions on the rung, which are included in the rung eigenenergies, thus raise the charge excitation energy by a substantial amount\cite{oldhanke}. For $t_\perp/t=1.0$, the charge response looks quite different. As shown in Fig.\ \ref{figcharge}(b) and (c), there is substantial spectral weight for both $q_\perp=0$ and $q_\perp=\pi$. For ${\bf q} \rightarrow (0,0)$, the density operator has even parity, causing the matrix element and thus the spectral weight to vanish, whereas at ${\bf q} \rightarrow (0,\pi)$, the density operator has odd parity so that optical transitions are allowed and there is spectral weight. In both channels, the size of the charge gap is approximately $4t$, and there is a broad structure of width $\sim 7t$. For $q_\perp=0$, most of the spectral weight occurs as a dispersive peak whose energy increases with increasing $q$, whereas for $q_\perp=\pi$ two peaks seem to contribute to the spectral weight distribution. The peaks are not well-defined enough over a range of $q$ to extract a dispersion, but the upper peak is heavily weighted near $q=\pi$, at $\omega \approx 9t$. We can qualitatively understand the broad incoherent structure of the charge response by considering particle--hole excitation in the single particle $A({\bf k},\omega)$ in Fig.\ \ref{figqmc1.0}. There are four dispersive bands and a broad background, so there should be weight in both the $q_\perp=0$ and $q_\perp=\pi$ branches of the charge response, and broad structure above a minimum excitation energy, which we see in Fig.\ \ref{figcharge}(c) and (d). From the single particle bands in Fig.\ \ref{figqmc0.5}, one can estimate the minimum particle--hole excitation energy to be $\sim 4t$. In Fig.\ \ref{figcharge}(c) and (d), the spectral weight near this minimum excitation energy is suppressed due to the particle--hole vertex. We have also carried out a calculation of charge response $\chi^{00}_{\text{RPA}}({\bf q},{\bf q}^\prime,\omega)$ within the antiferromagnetic RPA approximation described above, using the SDW dispersion in Eq.\ (\ref{eqAFHFdisp}), and also find a relatively broad structure above the charge gap for both $q_\perp=0$ and $q_\perp=\pi$. We plot the minimum excitation energy of $\chi^{00}_{\text{RPA}}({\bf q},{\bf q}^\prime,\omega)$ in Fig.\ \ref{figcharge}(b) and (c) as lines with solid dots. This line is located in the middle of the broad band in both plots. The spin and charge response functions for $t_\perp/t=0.5$ show the same general features as in the isotropic ($t_\perp/t=1.0$) case and are therefore not shown here. The spin susceptibility $\chi_s({\bf q},\omega)$ is also qualitatively identical to the RPA result $\chi^{+-}_{\rm RPA}({\bf q},{\bf q},\omega)$ with a smaller spin velocity than in the $t_\perp/t=1.0$ case. The charge susceptibility $\chi_c({\bf q},\omega)$ shows a clear dispersive band centered around the low--lying RPA excitations in the $q_\perp=0$ channel, whose energy increases with increasing $q$. In the $q_\perp=\pi$ channel, there is a broad structure of width $\sim 8t$ with again two peaks in the spectral weight distribution. \section{\label{CON} CONCLUSION} In summary, the single and two--particle dynamical properties of the two chain Hubbard model at half--filling can be understood by starting from two limits: the limit of non--interacting rungs treated exactly, which gives a good starting point for the large $t_\perp$ case for which the spin--spin correlation length along the chains is less than a lattice spacing, and the limit of an antiferromagnetically ordered state, which gives a good starting point for the small $t_\perp$ case when the spin--spin correlation length is large. The dynamical properties in the two regimes look quite different. For the large $t_\perp$ regime, the remnants of the level transitions of the two site system representing a single rung, suitably broadened into bands, can explain the major features of the single--particle, spin and charge responses. In the small $t_\perp$ case, calculations based on an antiferromagnetically ordered starting point, such as antiferromagnetic Hartree--Fock theory and spin--wave theory give a good qualitative picture of the coherent spin--wave part of the single particle spectral weight and of the two--particle spin dynamic correlation function. In addition, there is a broad incoherent band in the single particle spectral weight similar to that found in recent numerical work on the 1D and 2D systems. We also have shown the single particle spectral weight and the spin and charge response functions for the physical relevant, isotropic case ($t_\perp/t=1.0$). The results are qualitatively similar to these in the small $t_\perp$ region and can therefore be understood from an antiferromagnetic Hartree--Fock or spin--wave theory picture. The real ladder compounds have approximately the same coupling strength parallel to and perpendicular to the chains, and it will therefore be interesting to compare these isotropic $t_\perp/t=1.0$ results with future experiments. \section*{ACKNOWLEDGMENTS} We would like to thank R.\ Preuss, A.\ Muramatsu, and W.\ Ziegler for helpful discussions. H.E., W.H., and R.M.N.\ are grateful to the Bavarian "FORSUPRA" program on high $T_c$ research and the DFG under Grant No.\ Ha 1537/12-1 for financial support. D.P. acknowledges support from the EEC Human Capital and Mobility program under grant CHRX-CT93-0332, and D.J.S. from the NSF under Grant No.\ DMR 92-2507. The calculations were performed on Cray YMP's at the HLRZ in J\"ulich and at the LRZ M\"unchen. \newpage
\section{Introduction} \label{sect-introduction} In $3+1$ numerical relativity, one often wishes to locate the black hole(s) in a (spacelike) slice. As discussed by Refs.~\cite{Hawking-73 Hawking-Ellis}, a black hole is rigorously defined in terms of its event horizon, the boundary of future null infinity's causal past. Although the event horizon has, in the words of Hawking and Ellis (Ref.~\cite{Hawking-Ellis-quote}), ``a number of nice properties'', it's defined in an inherently {\em acausal\/} manner: it can only be determined if the entire future development of the slice is known. (As discussed by Refs.~\cite{ABBLMSSSW-94,LMSSW-95}, in practice the event horizon may be located to good accuracy given only the usual numerically generated approximate development to a nearly stationary state, but the fundamental acausality remains.) In contrast, an apparent horizon, also known as a marginally outer trapped surface, is defined (Refs.~\cite{Hawking-73,Hawking-Ellis}) locally in time, within a single slice, as a closed 2-surface whose outgoing null geodesics have zero expansion. An apparent horizon is slicing-dependent: if we define a ``world tube'' by taking the union of the apparent horizon(s) in each slice of a slicing, this world tube will vary from one slicing to another. In a stationary spacetime event and apparent horizons coincide, although this generally isn't the case in dynamic spacetimes. However, given certain technical assumptions, the existence of an apparent horizon in a slice implies the existence of an event horizon, and thus by definition a black hole, containing the apparent horizon. (Unfortunately, the converse doesn't always hold. Notably, Wald and Iyer (Ref.~\cite{Wald-Iyer-91}) have constructed a family of angularly anisotropic slices in Schwarzschild spacetime which approach arbitrarily close to $r = 0$ yet contain no apparent horizons.) There is thus considerable interest in numerical algorithms to find apparent horizons in numerically computed slices, both as diagnostic tools for locating black holes and studying their behavior (see, for example, Refs.~\cite{ABBLMSSSW-94,ABBHSS-94}), and for use ``on the fly'' during numerical evolutions to help in choosing the coordinates and ``steering'' the numerical evolution (Refs.~\cite{Thornburg-talk-91 Seidel-Suen-92,Thornburg-PhD,ADMSS-95}). This latter context makes particularly strong demands on a horizon-finding algorithm: Because the computed horizon position is used in the coordinate conditions, the horizon must be located quite accurately to ensure that spurious finite difference instabilities don't develop in the time evolution. Furthermore, the horizon must be re-located at each time step of the evolution, so the horizon-finding algorithm should be as efficient as possible. Finally, when evolving multiple-black-hole spacetimes in this manner it's desirable to have a means of detecting the appearance of a new outermost apparent horizon surrounding two black holes which are about to merge. We discuss this last problem further in section~\ref{sect-finding-outermost-apparent-horizons}. In this paper we give a detailed discussion of the \defn{Newton's method} apparent-horizon-finding algorithm. This algorithm poses the apparent horizon equation as a nonlinear elliptic (boundary-value) PDE on angular-coordinate space for the horizon shape function $r = h(\theta,\phi)$, finite differences this PDE, and uses some variant of Newton's method to solve the resulting set of simultaneous nonlinear algebraic equations for the values of $h$ at the angular-coordinate grid points. This algorithm is suitable for both axisymmetric and fully-general spacetimes, and we discuss both cases. As explained in section~\ref{sect-notation}, we assume a locally polar spherical topology for the coordinates and finite differencing, though we make no assumptions about the basis used in taking tensor components. \section{Notation} \label{sect-notation} Our notation generally follows that of Misner, Thorne, and Wheeler (Ref.~\cite{MTW}), with $G = c = 1$ units and a $(-,+,+,+)$ spacetime metric signature. We assume the usual Einstein summation convention for repeated indices regardless of their tensor character, and we use the Penrose abstract-index notation, as described by (for example) Ref.~\cite{Wald}. We use the standard $3+1$ formalism of Arnowitt, Deser, and Misner (Ref.~\cite{ADM-62}) (see Refs.~\cite{York-79 York-83} for recent reviews). We assume that a specific spacetime and $3+1$ (spacelike) slice are given, and all our discussions take place within this slice. We use the term \defn{horizon} to refer to the (an) apparent horizon in this slice. We often refer to various sets in the slice as being 1, 2, or 3-dimensional, meaning the number of {\em spatial\/} dimensions -- the time coordinate is never included in the dimensionality count. For example, we refer to the horizon itself as 2-dimensional. We assume that the spatial coordinates $x^i \equiv (r,\theta,\phi)$ are such that in some neighborhood of the horizon, surfaces of constant $r$ are topologically nested 2-spheres with $r$ increasing outward, and we refer to $r$ as a \defn{radial} coordinate and $\theta$ and $\phi$ as \defn{angular} coordinates. For pedagogical convenience (only), we take $\theta$ and $\phi$ to be the usual polar spherical coordinates, so that if spacetime is axisymmetric (spherically symmetric), $\phi$ is ($\theta$ and $\phi$ are) the symmetry coordinate(s). However, we make no assumptions about the detailed form of the coordinates, i.e.~we allow all components of the 3-metric to be nonzero. We emphasize that although our assumptions about the local topology of $r$ are fundamental, our assumptions about the angular coordinates are for pedagogical convenience only, and could easily be eliminated. In particular, all our discussions carry over unchanged to multiple black hole spacetimes, using (for example) either \v{C}ade\v{z}{} conformal-mapping equipotential coordinates (Ref.~\cite{Cadez-PhD}) or multiple-coordinate-patch coordinate systems (Ref.~\cite{Thornburg-87}). We use $ijkl$ for spatial (3-) indices, and $uvwxy$ for indices ranging over the angular coordinates only. $g_{ij}$ denotes the 3-metric in the slice, $g$ its determinant, and $\nabla_i$ the associated 3-covariant derivative operator. $K_{ij}$ denotes the 3-extrinsic curvature of the slice, and $K$ its trace. We use $\A$ to denote the 2-dimensional space of angular coordinates $(\theta,\phi)$. We sometimes need to distinguish between field variables defined on $\A$ or on the (2-dimensional) horizon, and field variables defined on a 3-dimensional neighborhood $\N$ of the horizon. This distinction is often clear from context, but where ambiguity might arise we use prefixes $\two$ and $\three$ respectively, as in $\two\! H$ and $\three\! H$. We use italic Roman letters $H$, $h$, etc., to denote {\em continuum\/} coordinates, functions, differential operators, and other quantities. We use sans serif Roman letters $\H$, $\h$, etc., to denote grid functions, and small capital Roman indices $\I$, $\J$, and $\K$ to index grid points. We use subscript grid-point indices to denote the evaluation of a continuum or grid function at a particular grid point, as in $H_\I$ or $\H_\I$. We use $\Jac[\P(\Q)]$ to denote the Jacobian matrix of the grid function $\P = \P(\Q)$, as defined by \eqref{eqn-Jac[P(Q)]}, and ${} \cdot {}$ to denote the product of two such Jacobians or that of a Jacobian and a grid function. We use $\Jacc[P(Q)]$ to denote the linearization of the differential operator $P = P(Q)$ about the point $Q$. We use $\MM$ as a generic finite difference molecule and $\M$ as a generic index for molecule coefficients. We write $\M \in \MM$ to mean that $\MM$ has a nonzero coefficient at position $\M$. Temporarily taking $\langle \M \rangle$ to denote some particular coordinate component of $\M$, we refer to $\max_{\M \in \MM} | \langle \M \rangle |$ as the \defn{radius} of $\MM$, and to the number of distinct $\langle \M \rangle$ values with $\M \in \MM$ as the \defn{diameter} or \defn{number of points} of $\MM$. (For example, the usual symmetric 2nd~order 3-point molecules for 1st~and 2nd~derivatives both have radius~1 and diameter~3.) We often refer to a molecule as itself being a discrete operator, the actual application to a grid function being implicit. Given a grid function $f$ and a set of points $\{ x_k \}$ in its domain, we use \,\,$\interp(f(x), x=a)$\,\, to mean an interpolation of the values $f(x_k)$ to the point $x=a$ and \,\,$\interp'(f(x), x=a)$\,\, to mean the derivative of the same interpolant at this point. More precisely, taking $I$ to be a smooth interpolating function (typically a Lagrange polynomial) such that $I(x_k) = f(x_k)$ for each $k$, \,\,$\interp(f(x), x=a)$\,\, denotes $I(a)$ and \,\,$\interp'(f(x), x=a)$\,\, denotes $\Bigl. (\partial I / \partial x) \Bigr|_{x=a}$. \section{The Apparent Horizon Equation} \label{sect-apparent-horizon-equation} As discussed by (for example) Ref.~\cite{York-89}, an apparent horizon satisfies the equation \begin{equation} H \equiv \two\! H \equiv \nabla_i n^i + K_{ij} n^i n^j - K = 0 \, \text{,} \label{eqn-horizon} \end{equation} where $n^i$ is the outward-pointing unit normal to the horizon, all the field variables are evaluated on the horizon surface, and where for future use we define the \defn{horizon function} $H \equiv \two\! H$ as the left hand side of~\eqref{eqn-horizon}. (Notice that in order for the 3-divergence $\nabla_i n^i$ to be meaningful, $n^i$ must be (smoothly) continued off the horizon, and extended to a field $\three n^i$ in some 3-dimensional neighborhood of the horizon. The off-horizon continuation is non-unique, but it's easy to see that this doesn't affect $H$ on the horizon.) To solve the apparent horizon equation~\eqref{eqn-horizon}, we begin by assuming that the horizon and coordinates are such that each radial coordinate line $\{ (\theta,\phi)=\text{constant} \}$ intersects the horizon in exactly one point. In other words, we assume that the horizon's coordinate shape is a \defn{\Strahlkorper}, defined by Minkowski as ``a region in $n$-dimensional Euclidean space containing the origin and whose surface, as seen from the origin, exhibits only one point in any direction'' (Ref.~\cite{Schroeder-quote}). Given this assumption, we can parameterize the horizon's shape by $r = h(\theta,\phi)$ for some single-valued \defn{horizon shape function} $h$ defined on the 2-dimensional domain $\A$ of angular coordinates $(\theta,\phi)$. Equivalently, we may write the horizon's shape as $\three\! F = 0$, where the scalar function $\three\! F$, defined on some 3-dimensional neighborhood $\N$ of the horizon, satisfies $\three\! F = 0$ if and only if $r = h(\theta,\phi)$, and we take $\three\! F$ to increase outward. In practice we take $\three\! F(r,\theta,\phi) = r - h(\theta,\phi)$. We define the non-unit outward-pointing normal (field) to the horizon by \begin{equation} s_i \equiv \three\! s_i = \nabla_i \three\! F \, \text{,} \end{equation} i.e.~by \begin{mathletters} \label{eqn-s-d(h)} \begin{eqnarray} s_r & = & 1 \\ s_u & = & - \partial_u h \, \text{.} \end{eqnarray} \end{mathletters} and the outward-pointing unit normal (field) to the horizon by \begin{eqnarray} n^i \equiv \three n^i & = & \frac{s^i}{\|s^k\|} \\ & = & \frac{g^{ij} s_j}{\sqrt{g^{kl} s_k s_l}} \label{eqn-n-u(s-d)} \\ & = & \frac{g^{ir} - g^{iu} \partial_u h} { \sqrt{ g^{rr} - 2 g^{ru} \partial_u h + g^{uv} (\partial_u h) (\partial_v h) } } \, \text{.} \label{eqn-n-u(h)} \end{eqnarray} Henceforth we drop the $\three$ prefixes on $\three\! s_i$ and $\three n^i$. Substituting \eqref{eqn-n-u(h)} into the apparent horizon equation~\eqref{eqn-horizon}, we see that the horizon function $H(h)$ depends on the (angular) 2nd derivatives of $h$. In fact, the apparent horizon equation~\eqref{eqn-horizon} is a 2nd~order elliptic (boundary-value) PDE for $h$ on the domain of angular coordinates $\A$. The apparent horizon equation~\eqref{eqn-horizon} must therefore be augmented with suitable boundary conditions to define a (locally) unique solution. These are easily obtained by requiring the horizon's 3-dimensional shape to be smooth across the artificial boundaries $\theta = 0$, $\theta = \pi$, $\phi = 0$, and $\phi = 2\pi$. \section{Algorithms for Solving the Apparent Horizon Equation} \label{sect-algorithms-survey} We now survey various algorithms for solving the apparent horizon equation~\eqref{eqn-horizon}. Ref.~\cite{Nakamura-Oohara-Kojima-87} reviews much of the previous work on this topic. In spherical symmetry, the apparent horizon equation~\eqref{eqn-horizon} degenerates into a 1-dimensional nonlinear algebraic equation for the horizon radius $h$. This is easily solved by zero-finding on the horizon function $H(h)$. This technique has been used by a number of authors, for example Refs.~\cite{Petrich-Shapiro-Teukolsky-85 Choptuik-PhD,Seidel-Suen-92,ADMSS-95}. (See also Ref.~\cite{Bizon-Malec-OMurchadha-88} for an interesting analytical study giving necessary and sufficient conditions for apparent horizons to form in non-vacuum spherically symmetric spacetimes.) In an axisymmetric spacetime, the angular-coordinate space $\A$ is effectively 1-di\-men\-sional, so the apparent horizon equation~\eqref{eqn-horizon} reduces to a nonlinear 2-point boundary value (ODE) problem for the function $h(\theta)$, which may be solved either with a shooting method, or with one of the more general methods described below. Shooting methods have been used by a number of authors, for example Refs.~\cite{Cadez-74,Dykema-PhD Abrahams-Evans-92,Bishop-82,Bishop-84,Shapiro-Teukolsky-92 Abrahams-Heiderich-Shapiro-Teukolsky-92}. The remaining apparent-horizon-finding algorithms we discuss are all applicable to either axisymmetric spacetimes (2-dimensional codes) or fully general spacetimes (3-dimensional codes). Tod (Ref.~\cite{Tod-91}) has proposed an interesting pair of \defn{curvature flow} methods for finding apparent horizons. Bernstein (Ref.~\cite{Bernstein-93}) has tested these methods in several axisymmetric spacetimes, and reports favorable results. Unfortunately, the theoretical justification for these methods' convergence is only valid in time-symmetric ($K_{ij} = 0$) slices. The next two algorithms we discuss are both based on a pseudospectral expansion of the horizon shape function $h(\theta,\phi)$ in some complete set of basis functions (typically spherical harmonics or symmetric trace-free tensors), using some finite number of the expansion coefficients $\{ a_k \}$ to parameterize of the horizon shape. One algorithm rewrites the apparent horizon equation $H(a_k) = 0$ as $\|H(a_k)\| = 0$, then uses a general-purpose function-minimization routine to search $\{ a_k \}$-space for a minimum of $\|H\|$. This algorithm has been used by Refs.~\cite{Brill-Lindquist-63,Eppley-77} in axisymmetric spacetimes, and more recently by Ref.~\cite{Libson-Masso-Seidel-Suen-95} in fully general spacetimes. Alternatively, Nakamura, Oohara, and Kojima (Refs.~\cite{Nakamura-Kojima-Oohara-84,Oohara-Nakamura-Kojima-85 Oohara-86}) have suggested a functional iteration method for directly solving the apparent horizon equation $H(a_k) = 0$ for the expansion coefficients $\{ a_k \}$, and have used it in a number of fully general spacetimes. Kemball and Bishop (Ref.~\cite{Kemball-Bishop-91}) have suggested and tested several modifications of this latter algorithm to improve its convergence properties. The final algorithm we discuss, and the main subject of this paper, poses the apparent horizon equation $H(h) = 0$ as a nonlinear elliptic (boundary-value) PDE for $h$ on the angular-coordinate space $\A$. Finite differencing this PDE on an angular-coordinate grid $\{ (\theta_\K,\phi_\K) \}$ gives a set of simultaneous nonlinear algebraic equations for the unknown values $\{ h(\theta_\K,\phi_\K) \}$, which are then solved by some variant of Newton's method. This \defn{Newton's-method} algorithm (we continue to use this term even if a modification of Newton's method is actually used) has been used in axisymmetric spacetimes by a number of authors, for example Refs.~\cite{Eardley-75,Cook-PhD,Cook-York-90,Cook-Abrahams-92 Thornburg-PhD}, and is also applicable in fully general spacetimes when the coordinates have a (locally) polar spherical topology. Huq (Ref.~\cite{Huq-talk-93}) has extended this algorithm to fully general spacetimes with Cartesian-topology coordinates and finite differencing, and much of our discussion remains applicable to his extension. The Newton's-method algorithm has three main parts: the computation of the discrete horizon function $\H(\h)$, the computation of the discrete horizon function's Jacobian matrix $\Jac[\H(\h)]$, and the solution of the simultaneous nonlinear algebraic equations $\H(\h) = 0$. We now discuss these in more detail. \section{Computing the Horizon Function} \label{sect-horizon-function} In this section we discuss the details of the computation of the discrete horizon function $\H(\h)$. More precisely, first fix an angular-coordinate grid $\{ (\theta_\K,\phi_\K) \}$. Then, given a \defn{trial horizon surface} $r = h(\theta,\phi)$, which need not actually be an apparent horizon, we define $\h(\theta,\phi)$ to be the discretization of $h(\theta,\phi)$ to the angular-coordinate grid, and we consider the computation of $\H(\h)$ on the discretized trial horizon surface, i.e.~at the points $\{ ( {r=\h(\theta_\K,\phi_\K)}, {\theta=\theta_\K}, {\phi=\phi_\K} ) \}$. The apparent horizon equation \eqref{eqn-horizon} defines $H \equiv \two\! H$ in terms of the field variables and their spatial derivatives on the trial horizon surface. However, these are typically known only at the (3-dimensional) grid points of the underlying $3+1$ code of which the horizon finder is a part. We therefore extend $\two\! H$ to some (3-dimensional) neighborhood $\N$ of the trial horizon surface, i.e.~we define an extended horizon function $\three\! H$ on $\N$, \begin{eqnarray} \three\! H & = & \nabla_i n^i + K_{ij} n^i n^j - K \\ & = & \partial_i n^i + (\partial_i \ln \sqrt{g}) n^i + K_{ij} n^i n^j - K \, \text{.} \label{eqn-3H(n-u)} \end{eqnarray} To compute $\two \H(\h)$ on the (discretized) trial horizon surface, we first compute $\three \H(\h)$ on the underlying $3+1$ code's (3-dimensional) grid points in $\N$, then radially interpolate these $\three \H$ values to the trial-horizon-surface position to obtain $\two \H(\h)$, \begin{mathletters} \label{eqn-2H(3H)} \begin{equation} \two \H(\theta,\phi) = \interp \Bigl( \three \H(r,\theta,\phi), r=\h(\theta,\phi) \Bigr) \, \text{,} \end{equation} or equivalently \begin{equation} \two \H_\I = \interp \Bigl( \three \H_{\langle r\I \rangle}, r=\h_\I \Bigr) \, \text{,} \end{equation} \end{mathletters} where $\I$ is an angular grid-point index and the $\langle r\I \rangle$ subscript denotes that the interpolation is done independently at each angular coordinate along the radial coordinate line $\{ \theta=\theta_\I, \phi=\phi_\I \}$. In practice any reasonable interpolation method should work well here: Refs.~\cite{Cook-York-90 Cook-Abrahams-92} report satisfactory results using a spline interpolant; in this work we use a Lagrange (polynomial) interpolant centered on the trial-horizon-surface position, also with satisfactory results. Neglecting the interpolation error, we can also write~\eqref{eqn-2H(3H)} in the form \begin{equation} \two \H(\theta,\phi) = \three \H(r=\h(\theta,\phi), \theta, \phi) \, \text{.} \label{eqn-2H(3H)-no-interp} \end{equation} We consider two basic types of methods for computing the extended horizon function $\three \H(\h)$: \begin{itemize} \item A \defn{2-stage} computation method uses two sequential numerical finite differencing stages, first explicitly computing $\s_i$ and/or $\n^i$ by numerically finite differencing $\h$, then computing $\three \H$ by numerically finite differencing $\s_i$ or $\n^i$. \item A \defn{1-stage} computation method uses only a single numerical (2nd) finite differencing stage, computing $\three \H$ directly in terms of $\h$'s 1st~and 2nd angular derivatives. \end{itemize} Figure~\ref{fig-H(h)-methods} illustrates this. To derive the detailed equations for these methods, we substitute \eqrefs{eqn-s-d(h)} and \eqrefs{eqn-n-u(s-d)} into \eqref{eqn-3H(n-u)}: \begin{eqnarray} \three\! H & = & \nabla_i n^i + K_{ij} n^i n^j - K \\ & = & \partial_i n^i + (\partial_i \ln \sqrt{g}) n^i + K_{ij} n^i n^j - K \\ & = & \partial_i \frac{g^{ij} s_j}{(g^{kl} s_k s_l)^{1/2}} + (\partial_i \ln \sqrt{g}) \frac{g^{ij} s_j}{(g^{kl} s_k s_l)^{1/2}} + \frac{K^{ij} s_i s_j}{g^{kl} s_k s_l} - K \\ & = & {} \frac{A}{D^{3/2}} + \frac{B}{D^{1/2}} + \frac{C}{D} - K \, \text{,} \label{eqn-3H(ABCD)} \end{eqnarray} where the subexpressions $A$, $B$, $C$, and $D$ are given by \begin{mathletters} \label{eqn-ABCD(s-d)} \begin{eqnarray} A & = & {} - (g^{ik} s_k) (g^{jl} s_l) \partial_i s_j - \tfrac{1}{2} (g^{ij} s_j) \Bigl[ (\partial_i g^{kl}) s_k s_l \Bigr] \\ B & = & (\partial_i g^{ij}) s_j + g^{ij} \partial_i s_j + (\partial_i \ln \sqrt{g}) (g^{ij} s_j) \\ C & = & K^{ij} s_i s_j \\ D & = & g^{ij} s_i s_j \, \text{,} \end{eqnarray} \end{mathletters} i.e. \begin{mathletters} \label{eqn-ABCD(h)} \begin{eqnarray} A & = & (g^{ur} - g^{uw} \partial_w h) (g^{vr} - g^{vw} \partial_w h) \partial_{uv} h \nonumber \\ & & \quad {} - \tfrac{1}{2} (g^{ir} - g^{iu} \partial_u h) \Bigl[ \partial_i g^{rr} - 2 (\partial_i g^{ru}) \partial_u h + (\partial_i g^{uv}) (\partial_u h) (\partial_v h) \Bigr] \\ B & = & \Bigl[ \partial_i g^{ir} - (\partial_i g^{iu}) \partial_u h \Bigr] - g^{uv} \partial_{uv} h + (\partial_i \ln \sqrt{g}) (g^{ir} - g^{iu} \partial_u h) \\ C & = & K^{rr} - 2 K^{ru} \partial_u h + K^{uv} (\partial_u h) (\partial_v h) \\ D & = & g^{rr} - 2 g^{ru} \partial_u h + g^{uv} (\partial_u h) (\partial_v h) \, \text{.} \end{eqnarray} \end{mathletters} Comparing the 1-stage and 2-stage methods, the 2-stage methods' equations are somewhat simpler, so these methods are somewhat easier to implement and somewhat cheaper (faster) to compute. However, for a proper comparison the cost of computing the horizon function must be considered in conjunction with the cost of computing the horizon function's Jacobian. Compared to the 1-stage method, the 2-stage methods double the effective radius of the net $\H(\h)$ finite differencing molecules, and thus have 2(4)~times as many nonzero off-diagonal Jacobian elements for a 2(3)-dimensional code. In practice the cost of computing these extra Jacobian elements for the 2-stage methods more than outweighs the slight cost saving in evaluating the horizon function. We discuss the relative costs of the different methods further in section~\ref{sect-methods-comparison}. \section{Computing the Jacobian} \label{sect-Jacobian} In this section we discuss the details of the computation of the Jacobian matrix $\Jac[\H(\h)]$ of the horizon function $\H(\h)$ on a given trial horizon surface. \subsection{Computing the Jacobian of a Generic Function $\P(\Q)$} We consider first the case of a generic function $P(Q)$ in $d$~dimensions, finite differenced using $N$-point molecules. We define the Jacobian matrix of the discrete $\P(\Q)$ function by \begin{mathletters} \label{eqn-Jac[P(Q)]} \begin{equation} \Jac[\P(\Q)]_{\I\J} = \frac{\partial \P_\I}{\partial \Q_\J} \, \text{,} \end{equation} or equivalently by the requirement that \begin{equation} \delta \P_\I \equiv \Bigl[ \P(\Q + \delta \Q) - \P(\Q) \Bigr]_\I = \Jac[\P(\Q)]_{\I\J} \cdot \delta \Q_\J \end{equation} \end{mathletters} for any infinitesimal perturbation $\delta \Q$ of $\Q$. We assume that $\P$ is actually a {\em local\/} grid function of $\Q$, so the Jacobian matrix is sparse. (For example, this would preclude the nonlocal 4th~order \defn{compact differencing} methods described by Refs.~\cite{Ciment-Leventhal-75,Hirsh-75}.) We assume that by exploiting the locality of the discrete $\P(\Q)$ function, any single $\P_\I$ can be computed in $O(1)$ time, independent of the grid size. \subsubsection{Computing Jacobians by Numerical Perturbation} \label{sect-numerical-perturbation} We consider two general methods for computing the Jacobian matrix $\Jac[\P(\Q)]$. The first of these is the \defn{numerical perturbation} method. This involves numerically perturbing $\Q$ and examining the resulting perturbation in $\P(\Q)$, \begin{equation} \Jac[\P(\Q)]_{\I\J} \approx \left[ \frac{\P(\Q + \mu \e^{(\J)}) - \P(\Q)}{\mu} \right]_\I \, \text{,} \label{eqn-Jac[P(Q)]-NP} \end{equation} where $\e^{(\J)}$ is a Kronecker-delta vector defined by \begin{equation} \left[ \e^{(\J)} \right]_\I = \left\{ \begin{array}{l@{\quad}l} 1 & \text{if $\I = \J$} \\ 0 & \text{otherwise} \end{array} \right. \, \text{,} \end{equation} and $\mu$ is a ``small'' perturbation amplitude. This computation of the Jacobian proceeds by columns: for each $\J$, $\Q_\J$ is perturbed, and the resulting perturbation in $\P(\Q)$ gives the $\J$th column of the Jacobian matrix. The perturbation amplitude $\mu$ should be chosen to balance the truncation error of the one-sided finite difference approximation \eqref{eqn-Jac[P(Q)]-NP} against the numerical loss of significance caused by subtracting the nearly equal quantities $\P(\Q + \mu \e^{(\J)})$ and $\P(\Q)$. Refs.~\cite{Curtis-Reid-74,Stoer-Bulirsch-80-quote} discuss the choice of $\mu$, and conclude that if $\P(\Q)$ can be evaluated with an accuracy of $\varepsilon$, then $\mu \approx \sqrt{\varepsilon}$ ``seems to work the best''. In practice the choice of $\mu$ isn't very critical for horizon finding. Values of $10^{-4}$ to $10^{-6}$ seem to work well, and the inaccuracies in the Jacobian matrix resulting from these values of $\mu$ don't seem to be a significant problem. This method of computing Jacobians requires no knowledge of the $\P(\Q)$ function's internal structure. In particular, the $\P(\Q)$ function may involve arbitrary nonlinear computations, including multiple sequential stages of finite differencing and/or interpolation. This method is thus directly applicable to the $\two \H(\h)$ computation. Assuming that $\P(\Q)$ is already known, computing $\Jac[\P(\Q)]$ by numerical perturbation requires a total of $N^d$ $\P_\I$ evaluations at each grid point, i.e.~it requires a perturbed-$\P_\I$ evaluation for each nonzero Jacobian element. \subsubsection{Computing Jacobians by Symbolic Differentiation} \label{sect-symbolic-differentiation} An alternate method of computing the Jacobian matrix $\Jac[\P(\Q)]$ is by \defn{symbolic differentiation}. This method makes explicit use of the finite differencing scheme used to compute the discrete $\P(\Q)$ function. Suppose first that the continuum $P(Q)$ function is a position-dependent local {\em linear\/} differential operator, discretely approximated by a position-dependent local finite difference molecule $\MM$, \begin{equation} \P_\I = \sum_{\M \in \MM(\I)} \MM(\I)_\M \Q_{\I+\M} \, \text{.} \label{eqn-P(Q)-mol} \end{equation} Differentiating this, we have \begin{equation} \Jac[\P(\Q)]_{\I\J} \equiv \frac{\partial \P_\I}{\partial \Q_\J} = \left\{ \begin{array}{l@{\quad}l} \MM(\I)_{\J-\I} & \text{if $\J-\I \in \MM(\I)$} \\ 0 & \text{otherwise} \end{array} \right. \, \text{,} \label{eqn-Jac[P(Q)]-from-mol} \end{equation} so that the molecule coefficients at each grid point give the corresponding row of the Jacobian matrix. More generally, suppose $P$ is a position-dependent local nonlinear algebraic function of $Q$ and some finite number of $Q$'s derivatives, say \begin{equation} P = P(Q, \partial_i Q, \partial_{ij} Q) \, \text{.} \label{eqn-P(Q)-and-derivs} \end{equation} Logically, the Jacobian matrix $\Jac[\P(\Q)]$ is defined (by \eqref{eqn-Jac[P(Q)]}) in terms of the linearization of the discrete (finite differenced) $\P(\Q)$ function. However, as illustrated in figure~\ref{fig-linearize-vs-FD}, if the discretization (the finite differencing scheme) commutes with the linearization, we can instead compute the Jacobian by first linearizing the continuum $P(Q)$ function, then finite differencing this (continuum) linearized function. (This method of computing the Jacobian is essentially just the ``Jacobian part'' of the Newton-Kantorovich algorithm for solving nonlinear elliptic PDEs.) That is, we first linearize the continuum $P(Q)$ function, \begin{eqnarray} \delta P & = & \frac{\partial P}{\partial Q} \, {\delta Q} + \frac{\partial P}{\partial (\partial_i Q)} \, \delta \partial_i Q + \frac{\partial P}{\partial (\partial_{ij} Q)} \, \delta \partial_{ij} Q \\ & = & \frac{\partial P}{\partial Q} \, {\delta Q} + \frac{\partial P}{\partial (\partial_i Q)} \, \partial_i \delta Q + \frac{\partial P}{\partial (\partial_{ij} Q)} \, \partial_{ij} \delta Q \, \text{.} \label{eqn-P(Q)-and-derivs-linearized} \end{eqnarray} We then view the linearized function $\delta P(\delta Q)$ as a linear differential operator, and discretely approximate it by the position-dependent finite difference molecule \begin{equation} \MM = \frac{\partial P}{\partial Q} \II + \frac{\partial P}{\partial (\partial_i Q)} \dd_i + \frac{\partial P}{\partial (\partial_{ij} Q)} \dd_{ij} \, \text{,} \label{eqn-mol-from-P(Q)-Jac-coeffs} \end{equation} where $\II$ is the identity molecule and $\dd_i$ and $\dd_{ij}$ are finite difference molecules discretely approximating $\partial_i$ and $\partial_{ij}$ respectively. Finally, we apply \eqref{eqn-Jac[P(Q)]-from-mol} to the molecule $\MM$ defined by \eqref{eqn-mol-from-P(Q)-Jac-coeffs} to obtain the desired Jacobian matrix $\Jac[\P(\Q)]$. In practice, there's no need to explicitly form the molecule $\MM$ -- the Jacobian matrix elements can easily be assembled directly from the known $\II$, $\dd_i$, and $\dd_{ij}$ molecule coefficients and the \defn{Jacobian coefficients} $\partial P / \partial Q$, $\partial P / \partial (\partial_i Q)$, and $\partial P / \partial (\partial_{ij} Q)$. Once these coefficients are known, the assembly of the actual Jacobian matrix elements is very cheap, requiring only a few arithmetic operations per matrix element to evaluate \eqrefs{eqn-mol-from-P(Q)-Jac-coeffs} and~\eqrefb{eqn-Jac[P(Q)]-from-mol}. The main cost of computing a Jacobian matrix by symbolic differentiation is thus the computation of the Jacobian coefficients themselves. Depending on the functional form of the $P(Q)$ function, there may be anywhere from 1 to 10 coefficients, although in practice these often have many common subexpressions. In other words, where the numerical perturbation method requires a $\P_\I$ evaluation per nonzero Jacobian {\em element\/}, the symbolic differentiation method requires the computation of ``a few'' Jacobian-coefficient subexpressions per Jacobian {\em row\/}. More precisely, suppose the computation of all the Jacobian coefficients at a single grid point is $J$ times as costly as a $\P_\I$ evaluation. Then the symbolic differentiation method is approximately $N^d/J$ times more efficient than the numerical perturbation method. \subsection{Semantics of the Horizon Function Jacobian} \label{sect-Jacobian-semantics} We now consider the detailed semantics of the horizon function Jacobian. We define the Jacobian of $\H(\h) \equiv \two \H(\h)$, $\Jac[\H(\h)] \equiv \Jac[\two \H(\h)]$, by \begin{mathletters} \label{eqn-Jac[2H(h)]} \begin{equation} \Jac[\two \H(\h)]_{\I\J} = \frac{d \, \two \H_\I}{d \h_\J} \, \text{,} \end{equation} (where $\I$ and $\J$ are angular grid-point indices), or equivalently by the requirement that \begin{equation} \delta \two \H_\I \equiv \Bigl[ \two \H(\h + \delta \h) - \two \H(\h) \Bigr]_\I = \Jac[\two \H(\h)]_{\I\J} \cdot \delta \h_\J \end{equation} \end{mathletters} for any infinitesimal perturbation $\delta \h$. Here $\I$ and $\J$ are both angular (2-dimensional) grid-point indices. Notice that this definition uses the {\em total\/} derivative $d \, \two \H / d \h$. This is because $\two \H(\h)$ is defined to always be evaluated {\em at the position $r = \h(\theta,\phi)$ of the trial horizon surface\/}, so the Jacobian $\Jac[\two \H(\h)]$ must take into account not only the direct change in $\two \H$ at a fixed position due to a perturbation in $\h$, but also the implicit change in $\two \H$ caused by the field-variable coefficients in $\two \H$ being evaluated at a perturbed position $r = \h(\theta,\phi)$. It's also useful to consider the Jacobian $\Jac[\three \H(\h)]$ of the extended horizon function $\three \H(\h)$, which we define analogously by \begin{mathletters} \label{eqn-Jac[3H(h)]} \begin{equation} \Jac[\three \H(\h)]_{\I\J} = \frac{\partial \, \three \H_\I}{\partial \h_\J} \, \text{,} \end{equation} or equivalently by the requirement that \begin{equation} \delta \, \three \H_\I \equiv \Bigl[ \three \H(\h + \delta \h) - \three \H(\h) \Bigr]_\I = \Jac[\three \H(\h)]_{\I\J} \cdot \delta \h_\J \end{equation} \end{mathletters} for any infinitesimal perturbation $\delta \h$. Here $\I$ is a 3-dimensional grid-point index for $\three \H$, while $\J$ is an (angular) 2-dimensional grid-point index for $\h$. In contrast to $\Jac[\two \H(\h)]$, this definition uses the {\em partial\/} derivative $\partial \, \three \H / \partial \h$. This is because we take $\three \H(\h)$ to be evaluated at a fixed position (a grid point in the neighborhood $\N$ of the trial horizon surface) {\em which doesn't change with perturbations in $\h$\/}, so $\Jac[\three \H(\h)]$ need only take into account the direct change in $\three \H$ at a fixed position due to a perturbation in $\h$. $\Jac[\three \H(\h)]$ thus has much simpler semantics than $\Jac[\two \H(\h)]$. We have found $\Jac[\three \H(\h)]$ very useful, both as an intermediate variable in the computation of $\Jac[\two \H(\h)]$ (described in the next section), and also conceptually, as an aid to {\em thinking\/} about the Jacobians. \subsection{Computing the Horizon Function Jacobian} \label{sect-computing-horizon-function-Jacobian} Table~\ref{tab-methods-comparison} (discussed further in section~\ref{sect-methods-comparison}) summarizes all the Jacobian-computation methods in this paper, which we now describe in detail. We tag each method with a shorthand \defn{code}, which gives the method's basic properties: whether it computes $\Jac[\two \H(\h)]$ directly or computes $\Jac[\three \H(\h)]$ as an intermediate step, whether it uses symbolic differentiation or numerical perturbation, and whether it uses a 1-stage or a 2-stage horizon function computation. The simplest methods for computing $\Jac[\two \H(\h)]$ are the \defn{2-dimensional} ones, which work directly with $\two \H(\h)$ in angular-coordinate space, without computing $\Jac[\three \H(\h)]$ as an intermediate step. Since $\two \H(\h)$ isn't given by a simple molecule operation of the form \eqref{eqn-P(Q)-mol}, symbolic differentiation isn't directly applicable here. However, numerical perturbation in angular-coordinate space is applicable, using either a 1-stage or a 2-stage method to compute $\two \H(\h)$. We refer to the resulting Jacobian computation methods as the \defn{2d.np.1s} and \defn{2d.np.2s} methods respectively. Our remaining methods for computing $\Jac[\two \H(\h)]$ are all \defn{3-dimensional} ones, which first explicitly compute $\Jac[\three \H(\h)]$, then compute $\Jac[\two \H(\h)]$ from this in the manner described below. If $\three \H(\h)$ is computed using the 1-stage method, i.e.~via \eqrefs{eqn-3H(ABCD)} and~\eqrefb{eqn-ABCD(h)}, then either numerical perturbation or symbolic differentiation may be used to compute $\Jac[\three \H(\h)]$. We refer to these as the \defn{3d.np.1s} and \defn{3d.sd.1s} methods respectively. The symbolic-differentiation Jacobian coefficients for the 3d.sd.1s method are tabulated in appendix~\ref{app-SD-Jac-coeffs}. Alternatively, if $\three \H(\h)$ is computed using a 2-stage method, then $\Jac[\three \H(\h)]$ may be computed either by the simple numerical perturbation of $\three \H(\h)$ (the \defn{3d.np.2s} method), or by separately computing the Jacobians of the individual stages and matrix-multiplying them together. For the latter case, either numerical perturbation or symbolic differentiation may be used to compute the individual-stage Jacobians, giving the \defn{3d.np2.2s} and \defn{3d.sd2.2s} methods respectively. The symbolic-differentiation Jacobian coefficients for the 3d.sd2.2s method are tabulated in appendix~\ref{app-SD-Jac-coeffs}. For any of the 3-dimensional methods, once $\Jac[\three \H(\h)]$ is known, we compute $\Jac[\two \H(\h)]$ as follows: \begin{eqnarray} \Jac[\two \H(\h)]_{\I\J} & \equiv & \frac{d \, \two \H_\I}{d \h_\J} \\ & = & \frac{d \, \two \H(\theta_\I, \phi_\I)} {d \h(\theta_\J, \phi_\J)} \\ & = & \frac{d \, \three \H(r=\h(\theta_\I,\phi_\I), \theta_\I, \phi_\I)} {d \h(\theta_\J, \phi_\J)} \qquad \text{(by~\eqref{eqn-2H(3H)-no-interp})} \\ & = & \left. \frac{\partial \, \three \H(r, \theta_\I, \phi_\I)} {\partial \h(\theta_\J, \phi_\J)} \right|_{r = \h(\theta_\I, \phi_\I)} + \left. \frac{\partial \, \three \H(r, \theta_\I, \phi_\I)} {r} \right|_{r = \h(\theta_\I, \phi_\I)} \\ & = & \interp \Bigl( \Jac[\three \H(\h)]_{{\langle r\I \rangle} \J}, r = \h_\I \Bigr) + \interp' \Bigl( \three \H_{\langle r\I \rangle}, r = \h_\I \Bigr) \, \text{,} \label{eqn-Jac[2H(h)]-from-Jac[3H(h)]} \end{eqnarray} where the $\langle r\I \rangle$ subscripts in \eqref{eqn-Jac[2H(h)]-from-Jac[3H(h)]} denote that the interpolations are done along the radial line $\{ \theta=\theta_\I, \phi=\phi_\I \}$, analogously to \eqref{eqn-2H(3H)}, and where we neglect the interpolation errors in~\eqref{eqn-Jac[2H(h)]-from-Jac[3H(h)]}. Notice that the \,\,$\interp'(\dots)$\,\, term in~\eqref{eqn-Jac[2H(h)]-from-Jac[3H(h)]} may be computed very cheaply using the same $\three \H$ data values used in computing $\two \H$, cf.~\eqref{eqn-2H(3H)}. (The number of $\three \H$ data points used in the radial interpolation at each angular grid position will probably have to be increased by one to retain the same order of accuracy in the \,\,$\interp'(\dots)$\,\, term in \eqref{eqn-Jac[2H(h)]-from-Jac[3H(h)]} as in the \,\,$\interp(\dots)$\,\, term.) It's thus easy to compute $\Jac[\two \H(\h)]$ once $\Jac[\three \H(\h)]$ is known. \subsection{Comparing the Methods} \label{sect-methods-comparison} Table~\ref{tab-methods-comparison} summarizes all the horizon-function and Jacobian computation methods described in sections~\ref{sect-horizon-function} and~\ref{sect-computing-horizon-function-Jacobian}. The table also shows which Jacobian matrices the methods use, the methods' measured relative CPU times in our axisymmetric-spacetime (2-dimensional) code (discussed further in appendix~\ref{app-code-details}), and our estimates of the methods' approximate implementation effort (programming complexity). As can be seen from the table, for our implementation the 3d.sd.1s method is by far the most efficient of the Jacobian computation methods, being about a factor of $5$ faster than any of the numerical perturbation methods. In fact, the computation of the Jacobian $\Jac[\two \H(\h)]$ by the 3d.sd.1s method is only $1.5$--$2$ times more expensive than the simple evaluation of the horizon function $\two \H(\h)$. The relative performance of the different methods will of course vary considerably from one implementation to another, and especially between axisymmetric-spacetime (2-dimensional) and fully-general-spacetime (3-dimensional) codes. However, counting the number of operations needed for each method shows that the 3d.sd.1s method should remain the fastest for any reasonable implementation. (We omit details of the counting in view of their length and lack of general interest.) Notably, the 3d.sd.1s method's relative advantage over the other methods should be approximately a factor of the molecule diameter {\em larger\/} for fully-general-spacetime (3-dimensional) codes than for axisymmetric-spacetime (2-dimensional) codes such as ours. Considering now the implementation efforts required by the various methods, in general we find that these depend more on which Jacobian matrices are involved than on how the Jacobians are computed: The 2-dimensional methods, involving only $\Jac[\two \H(\h)]$, are the easiest to implement, while the 3-dimensional methods involving (only) $\Jac[\two \H(\h)]$ and $\Jac[\three \H(\h)]$ are somewhat harder to implement. The 3-dimensional methods involving the individual-stage Jacobians $\Jac[\s_i(\h)]$, $\Jac[\n^i(\h)]$, $\Jac[\three \H(\s_i)]$, and/or $\Jac[\three \H(\n^i)]$ are considerably more difficult to implement, due to these Jacobians' more complicated sparsity patterns. All the Jacobian matrices are highly sparse, and for reasonable efficiency it's essential to exploit this sparsity in their storage and computation. We have done this in our code, and our CPU-time measurements and implementation-effort estimates all reflect this. We briefly describe our sparse-Jacobian storage scheme in appendix~\ref{app-sparse-Jacobian-storage}. This scheme is very efficient, but its programming is a significant fraction of the overall Jacobian implementation effort, especially for the individual-stage Jacobians. Comparing numerical perturbation and symbolic differentiation methods, we had previously suggested (Ref.~\cite{Thornburg-PhD-SD-Jac-comments}) that symbolic-differentiation Jacobian computations would be very difficult to implement, necessarily requiring substantial support from a (computer) symbolic computation system. Several colleagues have expressed similar opinions to us. We had also previous suggested (Ref.~\cite{Thornburg-PhD-SD-Jac-comments}) that due to the structure of the $H(h)$ function, a Jacobian-coefficient formalism of the type described in sections~\ref{sect-symbolic-differentiation} and~\ref{sect-computing-horizon-function-Jacobian} would not be valid for the horizon function, so symbolic differentiation methods would require explicitly differentiating the finite difference equations. These suggestions have proven to be incorrect: using the Jacobian-coefficient formalism described in sections~\ref{sect-symbolic-differentiation} and~\ref{sect-computing-horizon-function-Jacobian}, only the continuum equations need be differentiated, and this is easily done by hand. More generally, using this formalism we find the actual programming of the symbolic differentiation methods to be only moderately more difficult than that of the numerical perturbation methods. Some of the Jacobian coefficients tabulated in appendix~\ref{app-SD-Jac-coeffs} are fairly complicated, but no more so than many other computations in $3+1$ numerical relativity. In order to be confident of the correctness of any of the Jacobian-computation methods except the simple 2-dimensional numerical perturbation ones, we feel that it's highly desirable to program an independent method (which may be programmed for simplicity at the expense of efficiency) and make an end-to-end comparison of the resulting Jacobian matrices. (We have successfully done this for each of the Jacobian matrices computed by each of the methods listed in table~\ref{tab-methods-comparison}, and our implementation-effort estimates there include doing this.) If, and only if, the Jacobians agree to within the expected truncation error of the numerical-perturbation Jacobian approximation~\eqref{eqn-Jac[P(Q)]-NP}, then we can have a high degree of confidence that both calculations are correct. If they disagree, then we find the detailed pattern of which matrix elements differ to be a very useful debugging aid. Summarizing our comparisons, then, we find that the best Jacobian computation method is clearly the 3d.sd.1s one. It's much more efficient than any of the other methods, and still quite easy to implement. \section{Convergence Tests} \label{sect-convergence-tests} Before continuing our discussion of Newton's-method horizon finding, in this section we digress to consider the convergence of finite differencing computations to the continuum limit. As has been forcefully emphasized by Choptuik (Refs.~\cite{Choptuik-PhD,Choptuik-91,Choptuik-Goldwirth-Piran-92}), a careful comparison of a finite differencing code's numerical results at different grid resolutions can yield very stringent tests of the code's numerical performance and correctness. In particular, such \defn{convergence tests} can yield reliable numerical estimates of a code's {\em external\/} errors, i.e.~of the deviation of the code's results from those that would be obtained by exactly solving the continuum equations. With, and only with, such estimates available, we can safely draw inferences about solutions of the continuum equations from the code's (finite-resolution) numerical results. To apply this technique in the horizon-finding context, suppose first that the (a) true (continuum) apparent horizon position $h^\ast$ is known. For a convergence test in this case, we run the horizon finder twice, using a 1:2 ratio of grid resolutions. As discussed in detail by Ref.~\cite{Choptuik-91}, if the code's numerical errors are dominated by truncation errors from $n$th~order finite differencing, the numerically computed horizon positions $\h$ must satisfy \begin{mathletters} \label{eqn-h(dx:dx/2)} \begin{eqnarray} \h[\Delta x] & = & h^\ast + (\Delta x)^n f + O((\Delta x)^{n+2}) \label{eqn-h(dx:dx/2)-dx} \\ \h[\Delta x / 2] & = & h^\ast + (\Delta x / 2)^n f + O((\Delta x)^{n+2}) \label{eqn-h(dx:dx/2)-dx/2} \end{eqnarray} \end{mathletters} at each grid point, where $\h[\Delta x]$ denotes the numerically computed horizon position using grid resolution $\Delta x$, and $f$ is an $O(1)$ smooth function depending on various high order derivatives of $h^\ast$ and the field variables, but {\em not\/} on the grid resolution. (We're assuming centered finite differencing here in writing the higher order terms as $O((\Delta x)^{n+2})$, otherwise they would only be $O((\Delta x)^{n+1})$.) Neglecting the higher order terms, i.e.~in the limit of small $\Delta x$, we can eliminate $f$ to obtain a direct relationship between the code's errors at the two resolutions, \begin{equation} \frac{\h[\Delta x / 2] - h^\ast}{\h[\Delta x] - h^\ast} = \frac{1}{2^n} \, \text{,} \label{eqn-conv-test-dx:dx/2} \end{equation} which must hold at each grid point common to the two grids. To test how well any particular set of (finite-resolution) numerical results satisfies this convergence criterion, we plot a scatterplot of the high-resolution errors $\h[\Delta x / 2] - h^\ast$ against the low-resolution errors $\h[\Delta x] - h^\ast$ at the grid points common to the two grids. If, and given the arguments of Ref.~\cite{Choptuik-91}, in practice {\em only\/} if, the error expansions~\eqref{eqn-h(dx:dx/2)} are valid with the higher order error terms negligible, i.e.~if and only if the errors are indeed dominated by the expected $n$th~order finite difference truncation errors, then all the points in the scatterplot will fall on a line through the origin with slope $1/2^n$. Now suppose the true (continuum) apparent horizon position $h^\ast$ is unknown. For a convergence test in this case, we run the horizon finder three times, using a 1:2:4 ratio of grid resolutions. Analogously to the 2-grid case, we now have \begin{mathletters} \label{eqn-h(dx:dx/2:dx/4)} \begin{eqnarray} \h[\Delta x] & = & h^\ast + (\Delta x)^n f + O((\Delta x)^{n+2}) \\ \h[\Delta x / 2] & = & h^\ast + (\Delta x / 2)^n f + O((\Delta x)^{n+2}) \\ \h[\Delta x / 4] & = & h^\ast + (\Delta x / 4)^n f + O((\Delta x)^{n+2}) \, \text{,} \end{eqnarray} \end{mathletters} at each grid point, with $f$ again independent of the grid resolution. Again neglecting the higher order terms, we can eliminate both $f$ and $h^\ast$ to obtain the \defn{3-grid} convergence criterion \begin{equation} \frac{\h[\Delta x / 2] - \h[\Delta x / 4]}{\h[\Delta x] - \h[\Delta x / 2]} = \frac{1}{2^n} \label{eqn-conv-test-dx:dx/2:dx/4} \end{equation} which must hold at each grid point common to the three grids. We test this criterion using a scatterplot technique analogous to that for the 2-grid criterion~\eqref{eqn-conv-test-dx:dx/2}. We emphasize that for a 3-grid convergence test of this type, the true continuum solution $h^\ast$ need not be known. In fact, nothing in the derivation actually requires $h^\ast$ to be the true continuum horizon position -- it need only be the true continuum solution to some continuum equation such that the truncation error formulas \eqref{eqn-h(dx:dx/2:dx/4)} hold. We make use of this latter case in sections~\ref{sect-Newton-Kantorovich-method} and~\ref{sect-global-conv-HSF-errors} to apply 3-grid convergence tests to intermediate Newton iterates (trial horizon surfaces) of our horizon finder. For both the 2-grid and the 3-grid convergence test, we find that the {\em pointwise\/} nature of the scatterplot comparison makes it significantly more useful than a simple comparison of gridwise norms. In particular, the scatterplot comparison clearly shows convergence problems which may occur only in a small subset of the grid points (for example near a boundary), which would be ``washed out'' in a comparison of gridwise norms. Notice also that the parameter $n$, the order of the convergence, is (should be) known in advance from the form of the finite differencing scheme. Thus the slope-$1/2^n$ line with which the scatterplot points are compared isn't fitted to the data points, but is rather an a~priori prediction with {\em no\/} adjustable parameters. Convergence tests of this type are thus a very strong test of the validity of the finite differencing scheme and the error expansions \eqref{eqn-h(dx:dx/2)} or~\eqref{eqn-h(dx:dx/2:dx/4)}. \section{Solving the Nonlinear Algebraic Equations} \label{sect-nonlinear-algebraic-equations} Returning to our specific discussion of horizon finding, we now discuss the details of using Newton's method or a variant to solve the simultaneous nonlinear algebraic equations $\H(\h) = 0$. \subsection{Newton's Method} \label{sect-Newton's-method} The basic Newton's-method algorithm is well known: At each iteration, we first linearize the discrete $\H(\h)$ function about the current approximate solution $\h^{(k)}$, \begin{equation} \H(\h^{(k)} + \delta \h) = \H(\h^{(k)}) + \Jac[\H(\h^{(k)})] \cdot \delta \h + O(\|\delta \h\|^2) \, \text{,} \label{eqn-H(h)-linearized} \end{equation} where $\delta \h$ now denotes a finite perturbation in $\h$, and where $\Jac[\H(\h^{(k)})]$ denotes the Jacobian matrix $\Jac[\H(\h)]$ evaluated at the point $\h = \h^{(k)}$. We then neglect the higher order (nonlinear) terms and solve for the perturbation $\delta \h^{(k)}$ such that $\H(\h^{(k)} + \delta \h^{(k)}) = 0$. This gives the simultaneous linear algebraic equations \begin{equation} \Jac[\H(\h^{(k)})] \cdot \delta \h^{(k)} = - \H(\h^{(k)}) \label{eqn-Newton-delta-h} \end{equation} to be solved for $\delta \h^{(k)}$. Finally, we update the approximate solution via \begin{equation} \h^{(k+1)} \leftarrow \h^{(k)} + \delta \h^{(k)} \, \text{,} \label{eqn-Newton-h-update} \end{equation} and repeat the iteration until some convergence criterion is satisfied. Notice that here we're using the word ``convergence'' in a very different sense from that of section~\ref{sect-convergence-tests} -- here it refers to the \defn{iteration-convergence} of the Newton iterates $\h^{(k)}$ to the exact solution $\h^\ast$ of the discrete equations, whereas there it refers to the \defn{finite-difference-convergence} of a finite difference computation result $\h[\Delta x]$ to its continuum limit $h^\ast$ as the grid resolution is increased. Once the current solution estimate $\h^{(k)}$ is reasonably close to $\h^\ast$, i.e.~in practice once the trial horizon surface is reasonably close to the (an) apparent horizon, Newton's method converges extremely rapidly. In particular, once the linear approximation in \eqref{eqn-H(h)-linearized} is approximately valid, Newton's method roughly squares the relative error $\|\h - \h^\ast\| / \|\h^\ast\|$ at each iteration, and can thus bring the error down to a negligible value in only a few (more) iterations. (This rapid \defn{quadratic} convergence depends critically on the mutual consistency of the horizon function and Jacobian matrix used in the computation, and is thus a useful diagnostic for monitoring the Jacobian's correctness.) (For a detailed discussion of Newton's method, including precise formulations and proofs of these statements, see, for example, Ref.~\cite{Stoer-Bulirsch-80-Newton's-method}.) However, if the initial guess $\h^{(0)}$ for the horizon position, or more generally any Newton iterate (trial horizon surface) $\h^{(k)}$, differs sufficiently from $\h^\ast$ so that the linear approximation in \eqref{eqn-H(h)-linearized} isn't approximately valid, then Newton's method may converge poorly, or fail to converge at all. \subsection{Modifications of Newton's Method} \label{sect-modifications-of-Newton's-method} Unfortunately, as discussed in section~\ref{sect-global-conv-HSF-errors}, for certain types of initial guesses Newton's method fails to converge unless the initial guess is very close to the exact solution of the finite difference equations. There's an extensive numerical analysis literature on more robust \defn{modified Newton} algorithms for solving nonlinear algebraic equations, for example Refs.~\cite{Bank-Rose-80 Bank-Rose-81,Dennis-Schnabel-83,MINPACK,Numerical-Recipes-2nd-edition ZIB-90-11,ZIB-91-10}. We have found Ref.~\cite{Dennis-Schnabel-83} to be a particularly useful introduction to this topic. For horizon finding, the Jacobian matrix's size is the number of angular grid points on the horizon surface. This is generally large enough that it's important for the nonlinear-algebraic-equations solver to support treating the Jacobian as either a band matrix (for axisymmetric-spacetime codes) or a fully general sparse matrix (for fully-general-spacetime codes). It's also desirable for the nonlinear-algebraic-equation solver to permit explicit bounds on the solution vector, so as to ensure the trial horizon surfaces never fall outside the radial extent of the code's main 3-dimensional grid. Unfortunately, these requirements rule out many nonlinear-algebraic-equation software packages. For the sake of expediency, in the present work we chose to write our own implementation of a relatively simple modified-Newton algorithm, the \defn{line search} algorithm described by Refs.~\cite{Dennis-Schnabel-83,Numerical-Recipes-2nd-edition}. However, a much better long-term solution would be to use an extant nonlinear-algebraic-equations code embodying high-quality implementations of more sophisticated algorithms, such as the {\sc GIANT} code described by Refs.~\cite{ZIB-90-11,ZIB-91-10}. We would expect Newton's-method horizon-finding codes using such software to be considerably more robust and efficient than our present code. The modified-Newton algorithm used in this work, the line-search algorithm of Refs.~\cite{Dennis-Schnabel-83,Numerical-Recipes-2nd-edition}, is identical to the basic Newton's-method algorithm, except that the Newton's-method update \eqref{eqn-Newton-h-update} is modified to $\h^{(k+1)} \leftarrow \h^{(k)} + \lambda \, \delta \h^{(k)}$, where $\lambda \in (0,1]$ is chosen at each \defn{outer} iteration by an inner \defn{line search} iteration to ensure that $\|\H\|_2$ decreases monotonically. Refs.~\cite{Dennis-Schnabel-83 Numerical-Recipes-2nd-edition} show that such a choice of $\lambda$ is always possible, and describe an efficient algorithm for it. Sufficiently close to the solution $\h^\ast$, this algorithm always chooses $\lambda = 1$, and so takes the same steps as Newton's method. The overall modified-Newton algorithm thus retains the extremely rapid convergence of Newton's method once the linear approximation in \eqref{eqn-H(h)-linearized} is good. The line-search algorithm described by Refs.~\cite{Dennis-Schnabel-83 Numerical-Recipes-2nd-edition} always begins by trying the basic Newton step $\lambda = 1$. For horizon finding, we have slightly modified the algorithm to decrease the starting value of $\lambda$ if necessary to ensure that $\h^{(k)} + \lambda \, \delta \h^{(k)}$ lies within the radial extent of our code's main (3-dimensional) numerical grid at each angular grid coordinate. Our implementation of the algorithm also enforces an upper bound (typically 10\%) on the relative change $\|\lambda \, \delta \h^{(k)} / \h^{(k)}\|_\infty$ in any component of $\h^{(k)}$ in a single outer iteration. However, it's not clear whether or not this latter restriction is a good idea: although it makes the algorithm more robust when the $\H(\h)$ function is highly nonlinear, it may slow the algorithm's convergence when the $\H(\h)$ function is only weakly nonlinear and the error in the initial guess is large. We give an example of this latter behavior in section~\ref{sect-accuracy}. \subsection{The Newton-Kantorovich Method} \label{sect-Newton-Kantorovich-method} We have described the Newton's-method algorithm, and the more robust modified versions of it, in terms of solving the discrete $\H(\h) = 0$ equations. However, these algorithms can also be interpreted directly in terms of solving the continuum $H(h) = 0$ equations. This \defn{Newton-Kantorovich} method, and its relationship to the discrete Newton's method, are discussed in detail by Ref.~\cite{Boyd}. For the Newton-Kantorovich algorithm, at each iteration, we first linearize the continuum differential operator $H(h)$ about the current continuum approximate solution $h^{(k)}$, \begin{equation} H(h^{(k)} + \delta h) = H(h^{(k)}) + \Jacc[H(h^{(k)})] (\delta h) + O(\|\delta h\|^2) \, \text{,} \label{eqn-continuum-H(h)-linearized} \end{equation} where $\delta h$ is now a finite perturbation in $h$, and where the linear differential operator $\Jacc[H(h^{(k)})]$ is now the linearization of the differential operator $H(h)$ about the point $h = h^{(k)}$. We then neglect the higher order (nonlinear) terms and solve for the perturbation $\delta h^{(k)}$ such that $H(h^{(k)} + \delta h^{(k)}) = 0$. This gives the linear differential equation \begin{equation} \Jacc[H(h^{(k)})](\delta h^{(k)}) = - H(h^{(k)}) \label{eqn-Newton-Kantorovich-dh} \end{equation} to be solved for $\delta h^{(k)}$. Finally, we update the approximate solution via \begin{equation} h^{(k+1)} \leftarrow h^{(k)} + \delta h^{(k)} \, \text{,} \end{equation} and repeat the iteration until some convergence criterion is satisfied. Now suppose we discretely approximate this continuum Newton-Kantorovich algorithm by finite differencing the iteration equation \eqref{eqn-Newton-Kantorovich-dh}. If the finite differencing and the linearization commute in the manner discussed in section~\ref{sect-symbolic-differentiation}, then {\em this finite-difference approximation to the Newton-Kantorovich algorithm is in fact identical to the discrete Newton's-method algorithm applied to the (discrete) $\H(\h) = 0$ equations obtained by finite differencing the continuum $H(h) = 0$ equation\/}. (In a simpler context, our Jacobian-coefficient formalism described in section~\ref{sect-symbolic-differentiation} essentially just exploits the ``Jacobian part'' of this identity.) Therefore, when using the discrete Newton's method to solve the $\H(\h) = 0$ equations, we can equivalently view each Newton iterate (trial horizon surface) $\h^{(k)}[\Delta x]$ as being a finite difference approximation to the corresponding continuum Newton-Kantorovich iterate (trial horizon surface) $h^{(k)}$. As the grid resolution is increased, each Newton iterate $\h^{(k)}[\Delta x]$ should therefore show proper finite-difference-convergence {\em regardless of the iteration-convergence or iteration-divergence of the Newton or Newton-Kantorovich iteration itself\/}. Moreover, once we verify the individual Newton iterates' finite-differencing-convergence (with a 3-grid convergence test), we can safely extrapolate the iteration-convergence or iteration-divergence of this discrete iteration to that of the continuum Newton-Kantorovich algorithm applied to the (continuum) $H(h) = 0$ equations. In other words, by this procedure we can ascribe the iteration-convergence or iteration-divergence of Newton's method to inherent properties of the continuum $H(h) = 0$ equations, as opposed to (say) a finite differencing artifact. We make use of this in section~\ref{sect-global-conv-HSF-errors}. \section{Global Convergence of the Horizon Finder} \label{sect-global-convergence} We now consider the global convergence behavior of the Newton's-method horizon finding algorithm. That is, how close must the initial guess $\h^{(0)}$ be to the (an) exact solution $\h^\ast$ of the finite difference equations in order for the iterates (trial horizon surfaces) $\h^{(k)}$ to converge to $\h^\ast$? In other words, how large is the algorithm's radius of convergence? \subsection{Global Convergence for Schwarzschild Spacetime} To gain a general picture of the qualitative behavior of $H(h)$ and its implications for Newton's-method horizon finding, it's useful to consider Schwarzschild spacetime. We use the Eddington-Finkelstein slicing, where the time coordinate is defined by requiring $t + r$ to be an ingoing null coordinate. (These slices aren't maximal: $K$ is nonzero and spatially variable throughout the slices.) Taking the black hole to be of dimensionless unit mass, the (only) apparent horizon in such a slice is the coordinate sphere $r = 2$. More generally, a straightforward calculation gives \begin{equation} H = \frac{2 (r-2)}{r^{3/2} \sqrt{r + 2}} \end{equation} for spherical trial horizon surfaces with coordinate radius $r$. Figure~\ref{fig-Schw-H} shows $H(r)$ for these surfaces. As expected, $H = 0$ for the horizon $r = 2$. However, notice that $H$ reaches a maximum value at $r = r^{\max} = \tfrac{1}{2} (3 + \sqrt{33}) \approx 4.372$, and in particular that for $r > r^{\max}$, $H > 0$ and $dH / dr < 0$. Because of this, almost any algorithm -- including Newton's method and its variants -- which tries to solve $H(r) = 0$ using only local information about $H(r)$, and which maintains the spherical symmetry, will diverge towards infinity when started from within this region, or if any intermediate iterate (trial horizon surface) ever enters it. In fact, we expect broadly similar behavior for $H$ in any black hole spacetime: Given an asymptotically flat slice containing an apparent horizon or horizons, consider any 1-parameter family of topologically 2-spherical nested trial horizon surfaces starting at the outermost apparent horizon and extending outward towards the 2-sphere at spatial infinity. $H = 0$ for the horizon and for the 2-sphere at spatial infinity, so $\|H\|$ must attain a maximum for some finite trial horizon surface somewhere between these two surfaces. We thus expect the same general behavior as in the Schwarzschild-slice case, i.e.~divergence to infinity if the initial guess or any intermediate iterate (trial horizon surface) lies outside the maximum-$\|H\|$ surface. This argument isn't completely rigorous, since the algorithm could move inward in an angularly anisotropic manner, but this seems unlikely. Fortunately, in practice this isn't a problem: the black hole area theorem places an upper bound on the size of an apparent horizon, and this lets us avoid overly-large initial guesses, or restart the Newton iteration if any intermediate iterate (trial horizon surface) is too large. \subsection{Global Convergence in the Presence of High-Spatial-Frequency Errors} \label{sect-global-conv-HSF-errors} Assuming the initial guess is close enough to the horizon for the divergence-to-infinity phenomenon not to occur, we find the global convergence behavior of Newton's method to depend critically on the angular spatial frequency spectrum of the initial guess's error $\h^{(0)} - \h^\ast$: If the error has only low-spatial-frequency components (in a sense to be clarified below), then Newton's method has a large radius of convergence, i.e.~it will converge even for a rather inaccurate initial guess. However, {\em if the error has significant high-spatial-frequency components, then we find that Newton's method has a very small radius of convergence, i.e.~it often fails to converge even when the error $\h^{(0)} - \h^\ast$ is very small\/}. This behavior is {\em not\/} an artifact of insufficient resolution in the finite difference grid. Rather, it appears to be caused by a strong nonlinearity in the continuum $H(h)$ function for high-spatial-frequency components in $h$. In this context there's no sharp demarcation between ``low'' and ``high'' spatial frequencies, but in practice we use the terms to refer to angular Fourier components varying as (say) $\cos m\theta$ with $m \lesssim 4$ and $m \gtrsim 8$ respectively. \subsubsection{An Example} As an example of this behavior, consider Kerr spacetime with dimensionless angular momentum $a \equiv J/M^2 = 0.6$. We use the Kerr slicing, where the time coordinate is defined by requiring $t + r$ to be an ingoing null coordinate. (These slices generalize the Eddington-Finkelstein slices of Schwarzschild spacetime, and are similarly nonmaximal, with $K$ nonzero and spatially variable throughout the slices.) Taking the black hole to be of dimensionless unit mass, the (only) apparent horizon in such a slice is the coordinate sphere $r = h^\ast(\theta,\phi) = 1 + \sqrt{1 - a^2} = 1.8$. For this example we consider two different initial guesses for the horizon position: one containing only low-spatial-frequency errors, $r = h^{(0)}(\theta,\phi) = 1.8 + 0.1 \cos 4 \theta$, and one containing significant high-spatial-frequency errors, $r = h^{(0)}(\theta,\phi) = 1.8 + 0.1 \cos 10 \theta$. Notice that both initial guesses are quite close to the actual horizon shape, differing from it by slightly less than 5\%. We use a finite difference grid with $\Delta \theta = \frac{\pi/2}{50}$, which is ample to resolve all the trial horizon surfaces occurring in the example. Figure~\hbox{\ref{fig-Kerr-hp4-hp10}(a)} shows the behavior of Newton's method for the low-spatial-frequency-error initial guess. As can be seen, here Newton's method converges without difficulty. Figure~\hbox{\ref{fig-Kerr-hp4-hp10}(b)} shows the behavior of Newton's method for the high-spatial-frequency-error initial guess. Here Newton's method fails to converge: the successive iterates (trial horizon surfaces) $\h^{(k)}$ move farther and farther away the horizon, and rapidly become more and more nonspherical. Figure~\hbox{\ref{fig-Kerr-hp4-hp10}(c)} shows the behavior of the modified Newton's method for this same high-spatial-frequency-error initial guess. Although the first iteration still moves the trial horizon surface somewhat inward from the horizon, the nonsphericity damps rapidly, and the successive iterates (trial horizon surfaces) quickly converge to the horizon. Notice that all the intermediate iterates (trial horizon surfaces) in this example are well-resolved by the finite difference grid. To verify that insufficient grid resolution isn't a factor in the behavior of the horizon finder here, we have rerun all three parts of this example with several higher grid resolutions, obtaining results essentially identical to those plotted here. More quantitatively, following our discussion of the Newton-Kantorovich method in section~\ref{sect-Newton-Kantorovich-method}, we have made 3-grid convergence tests of each intermediate iterate (trial horizon surface) in this example. For example, figure~\ref{fig-Kerr-hp10-Newton-conv} shows a 3-grid convergence test for the Newton iterate (trial horizon surface) $\h^{(2)}$ plotted in figure~\hbox{\ref{fig-Kerr-hp4-hp10}(b)}, using grids with resolutions $\Delta \theta = \frac{\pi/2}{50}$:$\frac{\pi/2}{100}$:$\frac{\pi/2}{200}$. Notice that despite the iteration-divergence of the Newton iteration, this iterate shows excellent 4th~order finite-difference-convergence. The other Newton and modified-Newton iterates (trial horizon surfaces) in our example all similarly show excellent 4th~order finite-difference-convergence. We conclude that the iteration-divergence of Newton's method seen in figure~\hbox{\ref{fig-Kerr-hp4-hp10}(b)}, is in fact an inherent property of the continuum Newton-Kantorovich algorithm for this initial guess and slice. Looking at the internal structure of this algorithm, we see that its only approximation is the linearization of the continuum $H(h)$ function in \eqref{eqn-continuum-H(h)-linearized}, so the algorithm's iteration-divergence must (can only) be due to nonlinearity in the continuum $H(h)$ function. \subsubsection{The Horizon-Perturbation Survey} To investigate how general the poor convergence of Newton's method seen in this example is, and to what extent it also occurs for the modified Newton's method, we have made a Monte Carlo numerical survey of both algorithms' behavior over a range of different initial-guess-error spatial frequency spectra. For this survey we first fix a particular horizon-finding algorithm. Suppose we are given a slice containing an apparent horizon at the continuum position $h^\ast$, and consider running the horizon finder with the generic perturbed initial guess \begin{equation} h = h^\ast + \sum_{{\scriptstyle m = 0} \atop {\scriptstyle \text{$m$ even}}} ^M c_m \cos m\theta \end{equation} for some set of initial-guess-error Fourier coefficients $\{ c_m \}$. (Here we include only even-$m$ cosine terms in $\theta$ so as to preserve axisymmetry and equatorial reflection symmetry, which our code requires.) For each value of $M$ we define the horizon finder's \defn{convergence region} in $\{ c_m \}$-space to be the set of coefficients $\{ c_m \}$ for which the horizon finder converges (we presume to the correct solution). For example, the convergence region will in practice always include the origin in $\{ c_m \}$-space, since there $h = h^\ast$, so the initial guess differs from the exact solution of the discrete $\H(\h) = 0$ equations only by the small $\H(\h)$ finite differencing error. We define $V_M$ to be the (hyper)volume of the convergence region. As described in detail in appendix~\ref{app-hps-details}, we estimate $V_M$ by Monte Carlo sampling in $\{ c_m \}$-space. Given $V_M$, we then define the \defn{volume ratio} \begin{equation} R_M = \left\{ \begin{array}{l@{\quad}l} V_0 & \text{if $M = 0$} \\[1ex] {\displaystyle \frac{V_M}{V_{M-2}}} & \text{if $M \geq 2$} \end{array} \right. \, \text{,} \end{equation} so that $R_M$ measures the average radius of convergence of the horizon finder in the $c_M$ dimension. \subsubsection{Results and Discussion} We have carried out such a horizon-perturbation survey for the same Kerr slices of the unit-mass spin-$0.6$ Kerr spacetime used in the previous example, for both the Newton and the modified-Newton algorithms, for $M = 0$, $2$, $4$, \dots, $12$. Figure~\ref{fig-Kerr-hps} shows the resulting volume ratios. Although the precise values are somewhat dependent on the details of our implementation and on the test setup (in particular on the position of the inner grid boundary, which is at $r = 1$ for these tests), the relative trends in the data should be fairly generic. These tests use a grid with $\Delta \theta = \frac{\pi/2}{50}$, which is adequate to resolve all the perturbed trial horizon surfaces. As can be seen from the figure, the modified-Newton algorithm is clearly superior to the Newton algorithm, increasing the radius of convergence by a factor of $2$--$3$ at high spatial frequencies. However, both algorithms' radia of convergence still fall rapidly with increasing spatial frequency, approximately as $1 / M^{3/2}$, although the rate is slightly slower for the modified-Newton than for the Newton algorithm. The radius of convergence of Newton's method falls below~0.1 ($\sim \! 5\%$~of the horizon radius) by~$M \gtrsim 10$, and the data suggest that the radius of convergence of the modified-Newton method would be similarly small by~$M \gtrsim 18$. Since the grid resolution is adequate, we again conclude that the small radius of convergence of Newton's method must be due to a strong high-spatial-frequency nonlinearity in the continuum $H(h)$ function. Our horizon-perturbation survey covers only a single axisymmetric initial slice and generic axisymmetric perturbations of the initial guess, but it seems unlikely that the nonlinearity would diminish for more general cases. Huq (Ref.~\cite{Huq-95}) has made limited tests with nonaxisymmetric spacetimes and high-spatial-frequency perturbations, and has found (poor) convergence of Newton's method similar to our results. Although we write the continuum horizon function as $H = H(h)$, it's more accurate to write this as $H = H(g_{ij},K_{ij},h)$, since $H$ also depends on the slice's field variables and their spatial derivatives. Examining the functional form of the $H(g_{ij},K_{ij},h)$ function in~\eqrefs{eqn-3H(ABCD)} and~\eqrefb{eqn-ABCD(h)}, we see that $H$ depends on the $g^{ij}$ components in a manner broadly similar to its dependence on $h$. We thus conjecture that the $H(g_{ij},K_{ij},h)$ function may exhibit strong high-spatial-frequency nonlinearity in the field variables, in particular in the $g^{ij}$ components, similar to its nonlinear dependence on $h$. If this is the case, then high-spatial-frequency variations in the field variables, such as would be caused by high-frequency gravitational radiation, might well impair the convergence of Newton's method in a manner similar to high-spatial-frequency perturbations in $h$. Further investigation of this possibility, either by analytical study of the nonlinear structure of the $H(g_{ij},K_{ij},h)$ function, or by numerical investigations, would be very interesting. Fortunately, however, those (few) dynamic black hole spacetimes which have been explicitly computed thus far (for example Ref.~\cite{AHSSS-95}) seem to contain mainly low-frequency gravitational radiation. In general, how serious a problem is the poor high-spatial-frequency convergence of Newton's method? Given a sufficiently good initial guess, Newton's method still converges very rapidly (quadratically), so the key question is, how good is the initial guess in practice? Two cases seem to be of particular importance: If the horizon finder is being used to update a horizon's position at each time step of a time evolution, then the previous time step's horizon position probably provides a sufficiently good initial guess for Newton's method to converge well. In contrast, if the horizon finder is being used on initial data, or in a time evolution where there is no nearby horizon in the previous time step, then significant initial-guess errors can be expected, and Newton's method may converge poorly. \section{Accuracy of the Horizon Finder} \label{sect-accuracy} We now consider the accuracy of the Newton's-method horizon finding algorithm. That is, assuming the Newton or modified-Newton iteration converges, how close is the horizon finder's final numerically computed horizon position to the (a) true continuum horizon position $h^\ast$? The horizon finder computes Newton or modified-Newton iterates (trial horizon surfaces) $\h^{(k)}$ for $k = 0$, $1$, $2$, \dots, until some convergence criterion is satisfied, say at $k = p$. Because of the extremely rapid convergence of the Newton and modified-Newton iterations once the error is sufficiently small (cf.~section~\ref{sect-Newton's-method}), there's little extra cost in using a very strict convergence criterion, i.e.~in solving the discrete $\H(\h) = 0$ equations to very high accuracy. In our horizon finder we typically require $\| \H(\h^{(p)}) \|_\infty < 10^{-10}$. We denote the exact solution of the discrete $\H(\h) = 0$ equations by $\h^\ast$. Given that $\| \H(\h^{(p)}) \|$ is reasonably small, then from standard matrix-perturbation theory (see, for example, Refs.~\cite{Linpack-book-conditioning,Golub-Van-Loan-conditioning}), $\| \h^{(p)} - \h^\ast \| \lesssim \kappa \| \H(\h^{(p)}) \|$, where $\kappa$ is the condition number of the (presumably nonsingular) Jacobian matrix $\Jac[\H(\h)]$ at the horizon position. If we take the convergence tolerance to be strict enough for $\| \h^{(p)} - \h^\ast \|$ to be negligible, then the overall accuracy of the horizon finder, i.e.~the external error $\| \h^{(p)} - h^\ast \|$ in the computed horizon position, is thus limited only by the closeness with which the discrete $\H(\h) = 0$ equations approximate the continuum $H(h) = 0$ equations, i.e.~by the accuracy of the $\H(\h)$ finite differencing. This potential for very high accuracy is one of the main advantages of the Newton's-method horizon-finding algorithm. For an example of the accuracy attainable in practice, we again consider the Kerr slices of the unit-mass spin-$0.6$ Kerr spacetime. However, to make the horizon deviate from a coordinate sphere and hence be a more significant test case for our horizon finder, we apply the spatial coordinate transformation \begin{mathletters} \label{eqn-w4Kerr-coord-xform} \begin{equation} r \to r + \frac{b^2}{b^2 + r^2} \Bigl( a_2 \cos 2 \theta + a_4 \cos 4 \theta \Bigr) \end{equation} to the slice, where the parameters are given by \begin{equation} b = 5 \qquad a_2 = 0.75 \qquad a_4 = 0.05 \, \text{.} \end{equation} \end{mathletters} As shown in figure~\hbox{\ref{fig-w4Kerr}(a)}, in the transformed coordinates this gives a strongly non-spherical ``peanut-shaped'' horizon, similar in shape to those around a pair of coalescing black holes. We have run our horizon finder on this slice, using the warped-coordinate coordinate sphere $r = 1.8$ as an initial guess and a grid resolution of $\Delta \theta = \frac{\pi/2}{50}$. We used the modified-Newton algorithm, which converged to the horizon without difficulty. (The convergence took 9~iterations, but would have taken only 6~iterations in the absence of our 10\% restriction on the relative change in any component of $\h$ in a single outer iteration, cf.~section~\ref{sect-modifications-of-Newton's-method}.) Figure~\hbox{\ref{fig-w4Kerr}(a)} shows the initial guess and the final numerically computed horizon position. Figure~\hbox{\ref{fig-w4Kerr}(b)} shows the results of a 2-grid convergence test of the final numerically computed horizon position for this example, using grids with resolutions $\Delta \theta = \frac{\pi/2}{50}$:$\frac{\pi/2}{100}$. As can be seen, the numerically computed solution shows excellent 4th~order convergence. Moreover, the numerically computed horizon positions are very accurate, with $\| \h^{(p)} - h^\ast \| \sim 10^{-5} (10^{-6})$ for a grid resolution of $\Delta \theta = \frac{\pi/2}{50} (\frac{\pi/2}{100})$. Errors of this magnitude are typical of what we find for Newton's-method horizon finding using 4th~order finite differencing, so long as the grid adequately resolves the horizon shape. \section{Finding {\em Outermost\/} Apparent Horizons} \label{sect-finding-outermost-apparent-horizons} The main focus of this paper is on locally finding apparent horizons, i.e.~on finding an apparent horizon in a neighborhood of the initial guess. However, there's a related global problem of some interest which has heretofore attracted little attention, that of finding or recognizing the {\em outermost\/} apparent horizon in a slice. (By \defn{recognizing} the outermost apparent horizon we mean the problem of determining whether or not a given apparent horizon is in fact the outermost one in a slice.) These global problems are of particular interest when apparent horizons are used to set the inner boundary of a black-hole-excluding grid in the numerical evolution of a multiple-black-hole spacetime, as discussed by Refs.~\cite{Thornburg-talk-91,Seidel-Suen-92 Thornburg-PhD,ADMSS-95}. In this context, we can use the appearance of a new outermost apparent horizon surrounding the previously-outermost apparent horizons around two black holes as a diagnostic that the black holes have collided and coalesced into a single (distorted) black hole. As suggested by Ref.~\cite{Thornburg-PhD}, we can then generate a new numerical grid and attach it to the new outermost apparent horizon, and continue the evolution on the exterior of the new (distored) black hole. So far as we know, no reliable algorithms are known for finding or recognizing outermost apparent horizons in nonspherical spacetimes. (For spherical spacetimes, a 1-dimensional search on $H(r)$ suffices.) If started with a very large 2-sphere as the initial guess, the curvature flow method might well converge to the outermost horizon in the slice, but as mentioned in section~\ref{sect-algorithms-survey}, the theoretical justification for this method's convergence is only valid in time-symmetric ($K_{ij} = 0$) slices. For the remaining local-horizon-finding algorithms surveyed in section~\ref{sect-algorithms-survey}, including the Newton's-method one, we know of no better method for locating or recognizing outermost horizons than trying the local-horizon-finder with a number of different initial guesses near the suspected position of an outermost horizon. If this method succeeds it locates a horizon, but there's still no assurance that this horizon is the outermost one in the slice. Moreover, if all the local-horizon-finding trials fail, this may mean that there's no horizon in the vicinity of the initial guesses, or it may only mean that a horizon is present nearby but the method failed to converge to it. It's also not clear how many local-horizon-finding trials should be made, nor just how their initial guesses should be chosen. This is clearly not a satisfactory algorithm. Further research to develop reliable algorithms for finding or recognizing outermost apparent horizons in generic (nonspherical, nonmaximal) slices would be very useful. \section{Conclusions} We find Newton's method to be an excellent horizon-finding algorithm: it handles fully generic slices, it's fairly easy to implement, it's very efficient, it's generally robust in its convergence, and it's very accurate. These properties are all well known, and Newton's method is widely used for horizon finding. In this paper we focus on two key aspects of this algorithm: the computation of the Jacobian matrix, and the algorithm's global convergence behavior. Traditionally, the Newton's-method Jacobian matrix is computed by a numerical perturbation technique. In this paper we present a much more efficient ``symbolic differentiation'' technique. Conceptually, this entails differentiating the actual finite difference equations used to compute the discrete horizon function $\H(\h)$. However, provided the finite differencing scheme commutes with linearization, the computation can instead be done by first differentiating the continuum horizon function $H(h)$, then finite differencing. (This is essentially just the ``Jacobian part'' of the Newton-Kantorovich method for solving nonlinear PDEs.) In our axisymmetric-spacetime (2-dimensional) numerical code, this method is about a factor of~$5$ faster than than any other Jacobian computation method. In fact, the Jacobian computation using this method is only $1.5$--$2$ times more expensive than the simple evaluation of $\H(\h)$. We expect the symbolic differentiation method's relative advantage over other Jacobian computation methods to be roughly similar for other axisymmetric-spacetime (2-dimensional) codes, and an additional factor of $\sim \! 3$--$5$ larger for fully-general-spacetime (3-dimensional) codes. We had previously suggested (Ref.~\cite{Thornburg-PhD}) that symbolic-differentiation Jacobian computations would be quite difficult, necessarily requiring substantial support from a (computer) symbolic computation system. Several colleagues have expressed similar opinions to us. However, this turns out not to be the case: we computed all the symbolic-differentiation Jacobian coefficients for our horizon finder by hand in only a few few pages of algebra. Some of the coefficients are fairly complicated, but no more so than many other computations in $3+1$ numerical relativity. We find the actual programming of the symbolic differentiation Jacobian computation to be only moderately more difficult than that of a numerical perturbation computation. In order to be confident of the correctness of a symbolic differentiation Jacobian computation, we feel that it's highly desirable to program an independent numerical perturbation method and make an end-to-end comparison of the resulting Jacobian matrices. The comparison Jacobian computation may be programmed for simplicity at the expense of efficiency, so it needn't add much to the overall symbolic-differentiation implementation effort. Turning now to the convergence behavior of Newton's method, we find that so long as the error in the initial guess (its deviation from the true horizon position) contains only low-spatial-frequency components, a Newton's-method horizon finder has a large (good) radius of convergence, i.e.~it converges even for rather inaccurate initial guesses. However, if the error in the initial guess contains significant high-spatial-frequency components, then we find that Newton's method has a small (poor) radius of convergence, i.e.~it may fail to converge even when the initial guess is quite close to the true horizon position. In this context there's no sharp demarcation between ``low'' and ``high'' spatial frequencies, but in practice we use the terms to refer to angular Fourier components varying as (say) $\cos m\theta$ with $m \lesssim 4$ and $m \gtrsim 8$ respectively. Using a Monte Carlo survey of initial-guess-error Fourier-coefficient space, we find that the radius of convergence for Newton's method falls rapidly with increasing spatial frequency, approximately as $1 / m^{3/2}$. A simple ``line-search'' modification of Newton's method roughly doubles the horizon finder's radius of convergence, and slightly slows the rate of decline with spatial frequency. Using a robust nonlinear-algebraic-equations code to solve the discrete $\H(\h) = 0$ equations would probably give some further improvement, but we doubt that it would change the overall trend. Using quantitative convergence tests, we demonstrate that the poor high-spatial-frequency convergence behavior of Newton's method is {\em not\/} an artifact of insufficient resolution in the finite difference grid. Rather, it appears to be inherent in the (a) strong nonlinearity of the continuum $H(h)$ function for high-spatial-frequency components in $h$. We conjecture that $H$ may be similarly nonlinear in its high-spatial-frequency dependence on the inverse-metric components. If so, then the presence of high-frequency gravitational radiation might well also impair the convergence of Newton's method, and possibly other horizon-finding methods as well. Further investigation of this possibility would be very interesting. Fortunately, if the horizon finder is being used to update a horizon's position at each time step of a time evolution, then the previous time step's horizon position probably provides a sufficiently good initial guess for Newton's method to converge well. Provided it converges, the Newton's-method algorithm for horizon finding is potentially very accurate, in practice limited only by the accuracy of the $\H(\h)$ finite differencing scheme. Using 4th~order finite differencing, we demonstrate that the error in the numerically computed horizon position, i.e.~the deviation of $\h$ from the true continuum horizon position, shows the expected $O((\Delta \theta)^4)$ scaling with grid resolution $\Delta \theta$, and is typically $\sim \! 10^{-5} (10^{-6})$ for a grid resolution of $\Delta \theta = \frac{\pi/2}{50} (\frac{\pi/2}{100})$. Finally, we have argued that considerable further research is needed to develop algorithms for finding or recognizing the {\em outermost\/} apparent horizon in a slice. This is an important problem for the numerical evolution of multiple-black-hole spacetimes with the black holes excluded from the numerical evolution, but so far as we know no reliable algorithms are known for it except in spherical symmetry. \section*{Acknowledgments} We thank M.~Huq for numerous useful conversations on horizon finding, and for helpful comments on various drafts of this paper. We thank D.~Bernstein for communicating unpublished research notes on the curvature-flow method to us. We thank W.~G.~Unruh and the University of British Columbia Physics Department for their hospitality and the use of their research facilities. We thank J.~Wolfgang for major assistance in setting up computer facilities, and G.~Rodgers for financial support.
\section*{Figure Captions} \begin{description} \item {1.)} Comparison of the two structure functions used in the QCD analysis, MRS D-' and CTEQ2M, in the $x$ and $Q^2$ range relevant to the collider experiment, $Q^2 \sim (p_T^{\gamma})^2$. Also shown is the most recent parametrization MRS G. Fig.~1a: Density distribution of gluons, fig.~1b: Valence and sea quark distribution. \item {2a.)} Fractional difference between the NLO QCD calculation and data taken at low energies, $\sqrt{s} \leq 30.6$ GeV. \item {2b.)} Fractional difference between the NLO QCD calculation and data taken at the intermediate energy of $\sqrt{s}$ = 630 GeV by UA2. \item {2c.)} Fractional difference between the NLO QCD calculation and data taken at the Tevatron at $\sqrt{s}$ = 1.8 TeV. \end{description} \clearpage \end{document}
\section{#1}} \newcommand{\subsect}[1]{\subsection{#1}} \newcommand{\subsubsect}[1]{\subsubsection{#1}} \renewcommand{\theequation}{\thesection.\arabic{equation}} \renewcommand{\thefootnote}{\fnsymbol{footnote}} \def\footnote{\footnote} \footskip 1.0cm \def\sxn#1{\bigskip\bigskip \sect{#1} \medskip} \def\subsxn#1{\bigskip \subsect{#1} \medskip} \def\subsubsxn#1{\bigskip \subsubsect{#1} \medskip} \begin{document} \thispagestyle{empty} \setcounter{page}{0} \bibliographystyle{unsrt} \footskip 1.0cm \thispagestyle{empty} \setcounter{page}{0} \begin{flushright} SU-4240-615\\ July, 1995 \\ \end{flushright} \centerline {\bf A DISCRETIZED VERSION OF KALUZA-KLEIN THEORY } \centerline {\bf WITH TORSION AND MASSIVE FIELDS } \vspace*{10mm} \centerline {\bf Nguyen Ai Viet \footnote{ On leave of absence from the High Energy Division, Centre of Theoretical Physics, P.O.Box 429 Bo Ho 10000, Hanoi Vietnam.} \footnote{ Address after September 1, 1995: Physics Department, Rockerfeller University, New York, NY, USA } and Kameshwar C.Wali } \vspace*{3mm} \centerline {\it Department of Physics, Syracuse University,} \centerline {\it Syracuse, NY 13244-1130, U.S.A.} \vspace*{10mm} \normalsize \centerline {\bf Abstract} We consider an internal space of two discrete points in the fifth dimension of the Kaluza-Klein theory by using the formalism of noncommutative geometry developed in a previous paper \cite{VIWA} of a spacetime supplemented by two discrete points. With the nonvanishing internal torsion 2-form there are no constraints implied on the vielbeins. The theory contains a pair of tensor, a pair of vector and a pair of scalar fields. Using the generalized Cartan structure equation we are able not only to determine uniquely the hermitian and metric compatible connection 1-forms, but also the nonvanishing internal torsion 2-form in terms of vielbeins. The resulting action has a rich and complex structure, a particular feature being the existence of massive modes. Thus the nonvanishing internal torsion generates a Kaluza-Klein type model with zero and massive modes. \bigskip \newpage \section{ Introduction} It is generally believed that our current description of spacetime underlying both classical physics as well as quantum field theories is unsatisfactory and inadequate to deal with the description of phenomena at short distances. One is seeking a mathematical formalism that provides a quantum description of natural phenomena that, a priori, does not speak about spacetime in its basic formulation, but spacetime of classical physics as well as quantum field theories emerges in certain limiting regimes- just as classical behaviour of quantum systems can emerge in certain limiting regimes \cite{QUAN}. The recent proposal of Connes \cite{CO} and the so called noncommutative geometry (NCG) appears very promising towards the achievement of such a goal. It has given rise to the description of the Standard Model \cite{COLO} with new insights as regards spontaneous symmetry breaking and quark and lepton masses. It is natural to ask whether and how the classical general relativity fits into the scheme of NCG. The first step in answering this question was taken by Chamseddine et al \cite{CFF}, whose starting point was an abstract two-sheeted continuum that could be considered as the direct product of a single spacetime continuum and two discrete points. This led to gravity coupled to a Brans-Dicke scalar field. The scalar field can be interpreted as the distance between the two sheets \footnote{ More recently, other authors using different approaches have obtained essentially the same result. See Ref.\cite{VIWA} for references to related work .}. Similarly, it is always extremely tempting to give geometrical meaning to other physical fields. Thus, in the traditional Kaluza-Klein theory massless tensor, vector and scalar fields together with their massive excitations appear as result of extending the physical four-dimensional spacetime by an additional continuous fifth dimension. Unfortunately, the massive modes are infinite in number. In a previous paper \cite{VIWA}, we have developed the formalism for a discretized version of Kaluza-Klein theory within the framework of NCG. The starting point, as in Ref.\cite{CFF}, is an extended spacetime that includes two discrete points of the continuous internal fifth dimension of the Kaluza-Klein theory. We presented a generalization of the usual Riemannian geometry in the new context that demanded a vielbein consisting, to begin with, a pair of tensor, a pair of vector and a pair of scalar fields. Following the usual steps in building a theory of gravitation with the new geometry, we imposed torsion free, metric compatibility conditions on the connection 1-forms from which we constructed the action through the Ricci scalar curvature. We found that the imposed conditions altered the field content of the theory in a dramatic way, requiring in addition to the tensor, vector and scalar fields, new dilaton-like dynamical fields. The connection 1-forms and hence the Ricci scalar curvature were determined uniquely in terms of these fields. The resulting action provided a rich structure that lent itself to intriguing interpretations. One of the dilaton fields, for instance, could give rise to masses and cosmological constant. Moreover by imposing a reality condition on the vielbein 1-forms we could make the dilaton fields disappear leading to the zero-mode sector of the Kaluza-Klein theory as in Ref.\cite{LVW}. The previous NCG models that contain gravity coupled to the Brans-Dicke scalar can be considered as a particular case when the vector field is set to zero. While these interpretations are interesting in themselves to merit further study, we seek in this paper a formulation that does not alter the initial field content of the theory of two independent tensor, vector and scalar fields. From the viewpoint of the underlying mathematical framework of NCG, this is a reasonable requirement: the vielbein 1-forms should be free of constraints, retaining their most general form. The problem is how to achieve this. Now for physical reasons, it is necessary that we impose the metric compatibility condition. We recall that in the ordinary Cartan-Riemannian geometry the vanishing of torsion yields unique connection 1-forms in terms of the metric coefficients and their derivatives. Non-vanishing torsion requires additional information besides the metric. In our formulation, we find a way to avoid this situation. We impose a reality condition on the connection 1-forms and release the strict torsion free condition. In order to keep as close as possible to the usual Riemannian geometry, we assume that the usual spacetime indexed torsion 2-forms do vanish. However, we do not assume that the discrete internal space indexed torsion 2-form vanishes. This results in the unique determination of the related connection 1-forms. As we shall see, the nonvanishing internal torsion 2-form can be also determined in terms of the given vielbeins. This way we have an action that describes the general field content that we started with initially. The most remarkable result is that this discrete version of Kaluza-Klein theory contains a finite number of massive modes. The paper is organized as follows: In the next section we will review briefly the basic formalism and give the necessary formulas in order to make this paper self-contained. In Sect.3, we discuss how we compute the connection 1-forms, internal torsion and the Ricci scalar curvature. In Sect.4, we present the general structure of the action and consider special cases. The final section is devoted to a summary and conclusions. \setcounter{equation}{0} \section{ Two-point internal space and vielbein} \subsection{ Algebra of smooth functions and generalized derivatives} We consider a physical space-time manifold ${\cal M}$ extended by a discrete internal space of two points to which we assign a $Z_2$-algebraic structure. With this extended space-time, the customary algebra of smooth functions ${\cal C}^\infty ({\cal M})$ is generalized to ${\cal A} ~=~ {\cal C}^\infty({\cal M})\oplus{\cal C}^\infty({\cal M})$ and any generalized function $F \in {\cal A}$ can be written as \begin{equation} F(x)= f_+(x) e + f_-(x) r~ , \end{equation} where $ e,r \in Z_2 = \{ e,r~|~ e^2=e~,~ r^2=e~,~ er=re=r ~\}$. We adopt a $ 2 \times 2$ matrix representation for $ e,r $: \begin{equation} e =\pmatrix{1&0\cr 0&1\cr} ~~ ,~~ r=\pmatrix{1&0\cr 0&-1\cr}~. \end{equation} Then the function $ F(x)$ assumes a $2 \times 2$ matrix form, \begin{equation} F ~=~f_+(x)\pmatrix{1&0\cr 0&1\cr} ~ + ~ f_-(x) \pmatrix{1&0\cr 0&-1\cr} ~=~\pmatrix{f_1(x)&0\cr 0&f_2(x)\cr}~, \end{equation} where \begin{equation} f_\pm(x)= 1/2.( f_1(x) \pm f_2(x) ) . \end{equation} In this paper we will use small letters to denote the quantities of ordinary geometry and capital letters for generalized quantities of NCG. With the algebra ${\cal A}$ of smooth functions, we have what we may consider as the algebra of the generalized 0-forms $ \Omega^0({\cal M})= {\cal C}^\infty({\cal M}) \oplus {\cal C}^\infty({\cal M})$. To build the corresponding generalized higher forms, we need an exterior derivative or the Dirac operator $ D$ \cite{CO,COLO} in the language of NCG. For this purpose, as in Ref.\cite{VIWA}, let us define derivatives $ D_N ( N= \mu, 5) $ by \begin{eqnarray}\label{CODER} && D_\mu = \pmatrix{\partial_\mu &0\cr 0 &\partial_\mu\cr} ,~~~\mu = 0,1,2,3~, \cr && D_5 = \pmatrix{0& m\cr -m&0\cr} , \end{eqnarray} where $ m $ is a parameter with dimension of mass. We specify the action of the derivatives on the 0-form elements as given by \begin{equation}\label{COMDER} D_N(F)= [D_N, F] ~~,~~N=\mu, 5 ~~~~, \end{equation} satisfying the Newton-Leibnitz rule, \begin{equation} D_N(FG) = D_N(F) G + F D_N(G). \end{equation} Then the exterior derivative operator $D$ is given by \begin{eqnarray} D \doteq (~DX^\mu D_\mu~+~ DX^5 \sigma^\dagger D_5~), \end{eqnarray} where \begin{equation} \sigma^\dagger ~=~ \pmatrix{ ~0 & -1\cr 1 & 0 \cr}. \end{equation} $ DX^M $ are in general $ 2 \times 2 $ matrices that form a basis of the generalized 1-forms. They are direct generalizations of the differential elements. When spacetime becomes curved, as in general relativity (GR), $ DX^M $ denote a generalized curvi-linear differential elements. Their concrete form can be given in a concrete basis. The explicit form of $DX^M$ in the orthonormal basis will be given in the next subsection. \subsection{ General and orthonormal basis of 1-forms } The possible metric structure is guaranteed by the existence of a local orthonormal basis: the vielbein $ E^A $. Analogously to GR, if we work in the locally flat basis the vielbein $E^A$ can be chosen to be orthonormal. In Ref.\cite{VIWA}, we chose a diagonal representation for the curvi-linear basis $DX^\mu$ and $DX^5\sigma^\dagger$ to construct generalized one- and higher forms in analogy with the usual Riemannian geometry. However, it is more convenient to work in a representation in which the vielbeins $ E^A ( A= a,\dot 5) $ are diagonal. Locally, $E^A$ is given as follows \begin{eqnarray} E^a &~~=~~& \pmatrix{e^a & 0 \cr 0 & e^a \cr} ~~,\cr E^{\dot 5} &~~=~~& \pmatrix{ 0 & \theta \cr \theta & 0 \cr}~~, \end{eqnarray} where $ e^a $ is some ordinary vierbein 1-forms and $\theta $ is some hermitian Clifford element\footnote{ Completely, in analogy with GR, we can represent $ e^a $ and $ \theta $ as the locally flat Dirac matrices $\gamma^a $ and $ \gamma^5 $ as in the spinorial representation of Connes-Lott model.( This representation is used widely in literature. See for example \cite{CO,COLO,CFF} for details. However, in our formalism the two sheets are not necessarily the ones of different chiralities. Hence $ \theta $ in general will be kept as an abstract Clifford element).}. In this basis the wedge product can be defined as follows \begin{equation}\label{wedge} E^A \wedge E^B = - E^B \wedge E^A. \end{equation} In the orthonormal and locally flat basis $E^A$, the curvi-linear differential elements $ DX^M $ are in general not diagonal any more. Conversely, we can choose to work in the representation in which $ DX^M$ are diagonal. Then $E^A$ is not diagonal anymore as discussed in Ref.\cite{VIWA}. Both basis span the space of generalized 1-forms; hence an arbitrary 1-form $ U $ in NCG is given by \begin{equation} U~~=~~E^A U_A ~~= DX^M U_M , \end{equation} where $ U_A $ and $ U_M $ are the components of the 1-form $ U $ in the $E^A$ and $ DX^M $ basis respectively. As $E^A$ and $DX^M$ themselves are also 1-forms, we can express them in terms of each other as follows \begin{eqnarray} E^A~ &~~=~~& DX^M E^A_{~M} ~~,\cr DX^M &~~=~~& E^A E^M_{~A} ~~, \end{eqnarray} where $ E^A_M $ and $E^M_A$ are generalized functions satisfying \begin{eqnarray} E^A_{~N} E^N_{~B} &~~=~~& \delta ^A_{~B} \cr E^A_{~N} E^M_{~A} &~~=~~& \delta^M_{~N}~~ . \end{eqnarray} Without any loss of generality we can choose $ E^A_{~M} $ as follows : \begin{eqnarray}\label{VIELBEIN} E^a_\mu ~=~ \pmatrix{ e^a_{1\mu}(x) & 0 \cr 0 & e^a_{2\mu}(x) \cr}&~~,~~& E^a_5 ~~=~~0 \cr E^{\dot 5}_\mu ~~=~~ \pmatrix{ a_{1\mu}(x) & 0 \cr 0 & a_{2\mu}(x) \cr} = A_\mu &~~,~~& E^{\dot 5}_5 = \pmatrix{ \varphi_1(x) & 0 \cr 0 & \varphi_2(x) \cr} ~~=~~\Phi, \end{eqnarray} ( We use a ${\dot 5}$ index in the orthonormal basis to distinguish it from the index $5$ in the curvi-linear basis ). Thus \begin{equation}\label{TRANS} DX^\mu ~=~E^a E^\mu_a ~~,~~ DX^5~=(E^{\dot 5} - E^a A_a) \Phi^{-1}, \end{equation} where $ A_a = E^\mu_a A_\mu $. Now we can derive the transformation rules for the components of an arbitrary 1-form $ U $ between the two basis \begin{eqnarray} U_a ~~=~~ E^\mu_a ( U_\mu + A_\mu U_5) &~~~,~~~& U_{\dot 5} = \Phi^{-1} U_5 ~~, \cr U_\mu ~~=~~ E^a_\mu U_a - A_\mu \Phi U_{\dot 5} &~~~,~~~& U_5 = \Phi U_{\dot 5}~~. \end{eqnarray} To this end we note that the exterior derivative of a general 1-form $ U =DX^M U_M = E^A U_A $ is given by \begin{eqnarray} D U &~~=~~& (DX^\mu + DX^5\sigma^\dagger D_5 ) U ~~,\cr &~~=~~& E^a\wedge E^b ~( D U)_{ab} + E^a\wedge E^{\dot 5}~ 2 ({\cal D}U)_{a{\dot5}} ~~. \end{eqnarray} Using Eq.(\ref{TRANS}), we find \begin{eqnarray}\label{DER1F} ( DU)_{bc} &~~=~~& {1\over 2} E^\mu_{~b} E^\nu_{~c}( \partial_\mu E^a_{~\nu} - \partial_\nu E^a_{\mu}) U_a - {1\over 2} E^\mu_{~b} E^\nu_{~c} (\partial_\mu A_\nu -\partial_\nu A_\mu ) \Phi U_{\dot 5} \cr &~~~+~& {1\over 2} (E^\mu_{~b} \partial_\mu U_c - E^\nu_{~c} \partial_\nu U_b) + {m \over 2} [ (A_b {\tilde E}^\nu_{~c} -A_c {\tilde E}^\nu_{~b}) E^a_{~\nu} U_a + ( A_c {\tilde U}_b - A_b {\tilde U}_c) \cr &~~~+~& (A_b {\tilde E}^\nu_{~c} -A_c {\tilde E}^\nu_{~b}) ({\tilde A}_\nu -A_\nu) \Phi U_{\dot 5}] , \cr ( DU)_{b{\dot 5}} &~~=~~& {1\over 2}~{\tilde E}^\mu_{~b}~ ({\partial_\mu \Phi \over\Phi} U_{\dot 5} +\partial_\mu U_{\dot 5}) + {m\over 2}( \Phi^{-1}({\tilde U}_b -{\tilde E}^\mu_{~b}E^c_{~\mu}U_c) \cr &~~~+~& {\tilde E}^\mu_{~b}\Big( A_\mu - {\tilde A}_\mu) (1+ {\tilde \Phi}^{-1}\Phi)~\Big)U_{\dot 5} + {\tilde A}_b {\tilde U}_{\dot 5})~~ , \end{eqnarray} where we have redefined $ A_\mu $ in Eq.(\ref{VIELBEIN}) as $-A_\mu\Phi^{-1} $. In the $ E^A $ basis the hermitian conjugate of an arbitrary 1-form $ U = E^a U_a + E^{\dot 5}U_{\dot 5} $ is the 1-form $ U^\dagger = E^a U_a + E^{\dot 5} {\tilde U}_{\dot 5} $ where \begin{equation} {\tilde F} ~=~ \pmatrix{ f_2 & 0 \cr 0 & f_1\cr} ~~, ~~{\rm for~any~ function~} F ~=~\pmatrix{f_1 & 0 \cr 0 & f_2 \cr}~~. \end{equation} In the orthonormal basis, we have chosen $ E^A$ to be hermitian, the general 1-forms need not be hermitian, neither does the $DX^M$ basis. \subsection{Generalized metric} Following Ref.\cite{VIWA,LVW}, we define the metric ${\cal G}$ as the sesquilinear inner product of two 1-forms $ U $ and $ V$ satisfying \begin{eqnarray} \label{METRIC1} < U~ F~,~ V~ G > & ~=~ & F < U~,~V> G ~~ , \cr < U \otimes R ~,~ V\otimes S> &~=~& R^\dagger < U~,~V> S , \end{eqnarray} where $F, G$ are functions and $R, S $ are 1-forms. Assuming the existence of the local orthonormal basis $ E^A $, we have \begin{equation} \label{METRIC2} < E^A~, ~ E^B > ~~=~~ \eta^{AB}, \end{equation} where $\eta^{AB} ~=~ signature(~-~,~+~+~+~+~) $. {}From Eqs.(\ref{METRIC1}) and (\ref{METRIC2}) we obtain the generalized metric tensor in the familiar form \begin{eqnarray} {\cal G }_{M N } &~=~& E^{A}_{~M} \eta_{AB} E^B_{~N}, \cr {\cal G }^{M N} &~=~& E^M_{~A} \eta^{AB} E^N_{~B}. \end{eqnarray} With the vielbeins given in Eq.(\ref{VIELBEIN}), the components of the metric tensors ${\cal G}^{MN}$ and ${\cal G}_{MN}$ turn out to be \begin{eqnarray}\label{METRIC} {\cal G}^{\mu\nu} &~= ~& G^{\mu\nu} ~{\dot =}~ \pmatrix{ g_1^{\mu\nu} & 0 \cr 0 & g_2^{\mu\nu} \cr} , \cr {\cal G}^{\mu 5} &~ =~ & A^\mu ~=~ {\cal G}^{5 \mu} ~~ , \cr {\cal G}^{55} & = & \Phi^{-2}+ A^2 ~~,\cr {\cal G}_{\mu\nu} &~=~ &G_{\mu \nu} + A_\mu A_\nu ~{\dot =}~ \pmatrix{ g_{1\mu\nu} & 0 \cr 0 & g_{2\mu\nu}\cr} + A_\mu A_\nu ~~, \cr {\cal G}_{\mu 5} &~ =~ & {\cal G}_{5\mu} ~=~ A_{\mu}\Phi ~~,\cr G_{55} & = & \Phi^2 . \end{eqnarray} where $g_i^{\mu \nu} = e^\mu_{ia}\eta^{ab} e^\nu_{ib}~,~ i=1,2$ are the metric tensors on the two sheets. In passing we note that the components of the metric tensor are identical in form with those in the 5-dimensional Kaluza-Klein theory except that in the present case, the usual continuous $x^5$-dependence is replaced by the matrix form. \setcounter{equation}{0} \section{ Connection, torsion and curvature} \subsection {Hermitian and metric compatible connection 1-forms} We have shown in Ref.\cite{VIWA,LVW} that the metric compatible or Levi-Civita connection 1-form $\Omega_{AB}$ satisfies \footnote{ The interested reader might see Ref.\cite{VIWA} for the relation between the connection and the covariant derivative.} \begin{equation}\label{comp1} \Omega^{~~\dagger}_{~AB}~~=~~ -~ \Omega_{~BA}~~ . \end{equation} In this paper we will impose an additional reality condition on the connection, \begin{equation}\label{reality} \Omega^{~~\dagger}_{~AB} ~=~ \Omega_{~AB}~~. \end{equation} Together with the metric compatibility condition (\ref{comp1}) the reality condition implies the following conditions on the components $\Omega_{ABC}$ of the connection 1-form $\Omega_{AB}$ in the $E^A$ basis \begin{eqnarray}\label{comp2} \Omega_{~abc} & ~=~ & -~\Omega^{bac}, \cr \Omega_{~ab {\dot 5}} & ~=~ & -~\Omega_{~ba{\dot 5}}~ = ~ \omega_{~ab {\dot 5}} e~~,\cr \Omega_{~a{\dot 5}b} & ~=~ & - ~\Omega_{~{\dot 5}ab}, \cr \Omega_{~a{\dot 5}{\dot 5}} & ~=~ & -~ \Omega_{~{\dot 5}a{\dot 5}} ~=~ \omega_{~a{\dot 5}{\dot 5}} e,\cr \Omega_{~{\dot 5}{\dot 5}a}&~ =~ & \Omega_{~{\dot 5}{\dot 5}{\dot 5}} ~=~0 ~~. \end{eqnarray} That is to say, the reality condition (\ref{reality}) requires that the internal indexed components of the connection 1-forms are ordinary functions. \subsection{ The first structure equation and torsion 2-forms} The first Cartan structure equation defines the torsion 2-forms $ T^A $ as given by: \begin{equation}\label{TORSION} T^A = D E^A - E^B \wedge \Omega^A_{~~B}~, \end{equation} In Ref.\cite{VIWA}, we had assumed $ T^A = 0 ~~(A=a,{\dot 5})$ to determine the connection $\Omega$. As noted before, we were lead to a theory with a tensor, vector and scalar fields and also additional dilaton-like fields. In the present paper, we shall assume \begin{eqnarray}\label{TORFREE} T_{abc} &~~=~~& T_{ab{\dot 5}} ~~=~~0 ~~,\cr T_{{\dot 5}AB} &~~=~~& t_{{\dot 5}AB} r . \end{eqnarray} In other words, the torsion 2-forms involving the external physical spacetime index vanish while the torsion 2-form involving the internal index ${\dot 5}$ as in Eq.(\ref{TORFREE}) does not vanish. Then we can determine $ t^{\dot 5}_{~AB}$ as well as the hermitian and metric compatible connection 1-forms $ \Omega_{~AB}$ in terms of the vielbeins. Using the general formula (\ref{DER1F}), it is straightforward to compute the exterior derivatives $ DE^A$ needed to calculate $\Omega_{ABC}$ in Eq.(\ref{TORSION}). We omit the details and give only the results. \begin{eqnarray}\label{DERVIEL} (DE_a)_{bc} &~~=~~& -~( DE_a)_{cb}~~=~~ {1\over 2}~\Big [~ (E^\mu_{~b} E^\nu_{~c} - E^\mu_{~c}E^\nu_{~b})\partial_\mu E_{a\nu} ~~\cr &~~~+~& m \Big(~( A_b {\tilde E}^\nu_{~c} - A_c{\tilde E}^\nu_{~b})E_{a\nu} + (A_c\eta_{ab} - A_b\eta_{ac})~\Big)~\Big ]~~,\cr ( DE_a)_{b{\dot 5}} &~~=~~& -~( DE_a)_{{\dot 5}b}~~=~~ {m\over 2} \Phi^{-1}( \eta_{ab} - {\tilde E}^\mu_b E_a^{~ \mu} )~~, \cr ( DE_{\dot 5})_{bc} &~~=~~& -~( DE_{\dot 5})_{cb} {}~~=~~-~{1\over 2}~\Big [~(E^\mu_{~b} E^\nu_{~c} - E^\mu_{~c}E^\nu_{~b})\Phi \partial_\mu A_\nu ~~\cr &~~~+~& m (A_b {\tilde E}^\nu_{~c} - A_c{\tilde E}^\nu_{~b})({\tilde A}_\nu - A_\nu )\Phi {}~\Big]~~,\cr ( DE_{\dot 5})_{b{\dot 5}} &~~=~~& -~( DE_{\dot 5})_{{\dot 5}b}~~=~~ {1\over 2}~{\tilde E}^\mu_b \Big [~ {\partial_\mu \Phi \over \Phi } + m (A_\mu -{\tilde A}_\mu \Phi {\tilde \Phi }^{-1})~ \Big ] ~~. \end{eqnarray} In component form, the first Cartan structure equation reduces to \begin{equation} T_{ABC} ~~=~~ (DE_A)_{BC} - {1\over 2 }( \Omega_{ABC} - \Omega_{ACB})~~. \end{equation} With the condition (\ref{TORFREE}) on the torsion 2-forms we obtain \begin{equation} \Omega_{abc} ~~=~~ ( DE_a)_{bc}+( DE_b)_{ca}- ( DE_c)_{ab} {}~~, \end {equation} which in conjunction with Eqs.(\ref{DERVIEL}) determines $ \Omega_{abc}$ in terms of vielbeins. The condition (\ref{TORFREE}) together with Eq.(\ref{comp2}) leads to the following equation \begin{equation} \Omega_{ab{\dot 5}} - T_{{\dot 5} ab}~~ = ~~( DE_a)_{b{\dot 5}} + ( DE_b)_{{\dot 5}a}- ( DE_{\dot 5})_{ab} {}~~, \end{equation} from which we can determine $ \Omega_{ab{\dot 5}} = \omega_{ab{\dot 5}}e $ and $ T_{{\dot 5}ab} = t_{{\dot 5}ab} r $ . Finally, from \begin{equation} \Omega_{{\dot 5}c {\dot 5}} = 2 ( T_{{\dot 5} {\dot 5} c} - ( DE_{\dot 5} )_{{\dot 5} c}) \end{equation} we can determine $\Omega_{{\dot 5}c {\dot 5}}= \omega_{{\dot 5}c{\dot 5}} e $ and $T_{{\dot 5}{\dot 5} c} = t_{{\dot 5}{\dot 5} c} r$ . The final results are as follows : The components of the torsion 2-form $ T^{{\dot 5}}$ are \begin{eqnarray}\label{TOR5} T_{{\dot 5} ab} &~~=~~& {1\over 2}~ {\tilde E}^\mu_{~a}~\Big[~ {\tilde E}^\nu_{~b} {\tilde \Phi} {\tilde F}_{\mu\nu} - E^\nu_{~b}\Phi F_{\mu \nu } {}~\Big] + {m \over 4}~ \Big[ ~ (E^\mu_{~a} {\tilde E}_{b\mu } - E^\mu_{~b} {\tilde E}_{a\mu}) {\tilde \Phi}^{-1} \cr &~~~+~& \Phi^{-1}( {\tilde E}^\mu_{~a\mu } - {\tilde E}^\mu_a E_{b\mu }) + 2( {\tilde A}_\mu \Big(~(A_b {\tilde E}^\mu_{~b} - A_b {\tilde E}^\mu_{~a}) \Phi - ( {\tilde A}_a E^\mu_{~b} - A_b E^\mu_{~a}) {\tilde \Phi}~\Big] ~~,\cr T_{{\dot 5} a {\dot 5}} &~~=~~& {1\over 4} ~\Big [~ (~ {\tilde E}^\mu_{~b}{ \partial_\mu \Phi \over \Phi} - E^\mu_{~b}{\partial_\mu {\tilde \Phi} \over {\tilde \Phi}}~) + m (~ {\tilde E}^\mu_{~b} A_\mu - E^\mu_{~b} {\tilde A}_\mu \cr &~~~+~& A_b {\tilde \Phi} \Phi^{-1} - {\tilde A}_b \Phi {\tilde \Phi}^{-1}~)~\Big] ~~. \end{eqnarray} The components of the connection 1-forms $ \Omega_{AB} $ are given by: \begin{eqnarray}\label{OMEGA} \Omega_{abc} &~~=~~& {1\over 2}~\Big[~E^\mu_{~b}E^\nu_{~c} (\partial_\mu E_{a\nu} -\partial_\nu E_{a\mu }) + E^\mu_{~c}E^\nu_{~a} (\partial_\mu E_{b\nu} -\partial_\nu E_{b\mu }) - E^\mu_{~a}E^\nu_{~b} (\partial_\mu E_{c\nu} -\partial_\nu E_{c\mu })~\Big] \cr &~~~+~& {m \over 2}~ \Big[~ (A_b {\tilde E}^\nu_{~c} - A_c {\tilde E}^\nu_{~c} ) E_{a\nu} + (A_c {\tilde E}^\nu_{~a} - A_a {\tilde E}^\nu_{~c} ) E_{b\nu} - (A_a {\tilde E}^\nu_{~b} - A_b {\tilde E}^\nu_{~a} ) E_{c\nu} \cr &~~~+~& 2( A_a\eta_{cb} -A_b \eta_{ac})~ \Big] ~~,\cr \Omega_{ab{\dot 5}}&~~=~~& {1\over 4}~ \Big (~E^\mu_{~a} E^\nu_{~b} F_{\mu \nu} \Phi + {\tilde E}^\mu_{~a} {\tilde E}^\nu_{~b} {\tilde F}_{\mu \nu}{\tilde \Phi}~\Big) + {m \over 4}~ \Big[~ \Phi^{-1} ({\tilde E}^\mu_{~a} E_{b\mu} - {\tilde E}^\mu_{~b} E_{a\mu}) \cr &~~~+~& { \tilde \Phi}^{-1} ( E^\mu_{~a}{\tilde E}_{b\mu} - E^\mu_{~b} {\tilde E}_{a\mu}) + ({\tilde A}_\nu - A_\nu ) \Big(~( {\tilde A}_a E^\nu_{~b} - {\tilde A}_b E^\nu_{~a}){\tilde \Phi} \cr &~~~-~& ( A_a {\tilde E}^\nu_{~b} - A_b {\tilde E}^\nu_{~a}) \Phi~ \Big)~ \Big ] \cr \Omega_{{\dot 5}ab}&~~~=~& -~{1\over 4}~\Big(~E^\mu_{~a} E^\nu_{~b} F_{\mu \nu} \Phi + {\tilde E}^\mu_{~a} {\tilde E}^\nu_{~b} {\tilde F}_{\mu \nu}{\tilde \Phi}~\Big) + {m \over 4}~ \Big[~ \Phi^{-1}\Big(~ ( 4\eta_{ab} - (3 {\tilde E}^\mu_{~b} E_{a\mu} + {\tilde E}^\mu_{~a} E_{b\mu})~\Big) \cr &~~~+~& { \tilde \Phi}^{-1} ( E^\mu_{~b}{\tilde E}_{a\mu} + E^\mu_{~a} {\tilde E}_{b\mu}) - ({\tilde A}_\nu - A_\nu ) \Big(~ ( {\tilde A}_a E^\nu_{~b} - {\tilde A}_b E^\nu_{~a}){\tilde \Phi} \cr &~~~-~& ( A_a {\tilde E}^\nu_{~b} - A_b {\tilde E}^\nu_{~a}) \Phi~ \Big)~ \Big] \cr \Omega_{{\dot 5} b{\dot 5}} & ~~=~~& -~ \Omega_{b {\dot 5}{\dot 5}} ~~=~~ {1\over 2}~\Big[~{\tilde E}^\mu_b {\partial_\mu \Phi \over \Phi } + E^\mu_b {\partial_\mu {\tilde \Phi} \over {\tilde \Phi}}~ \Big] + { m \over 2}\Big[~ {\tilde E}^\mu_b ( A_\mu -{\tilde A}_\mu \Phi {\tilde \Phi }^{-1}) \cr &~~~+~& E^\mu_b ({\tilde A}_\mu - A_\mu {\tilde \Phi} \Phi^{-1})~\Big ] ~~. \end{eqnarray} \subsection { Second structure equation, curvature and the action} The second Cartan structure equation defines curvature 2-forms as follows \begin{equation}\label{CARTAN2} R_{AB} ~~=~~ D\Omega_{AB} + \Omega_{AC} \wedge \Omega^C_{~B} \end{equation} It is straightforward to use the expressions for the connection 1-forms given in Eq.(\ref{OMEGA}) to compute the components $ R_{ABCD}$ of the curvature 2-forms. We recall from Ref.\cite{VIWA} the expression of the Ricci scalar curvature \footnote{ The interested reader can see the Ref.\cite{VIWA} for the expression of R in an inner product form.} \begin{equation} R~~=~~ \eta^{AC} R_{ABCD} \eta^{BD}. \end{equation} After a lengthy but straightforward calculation we obtain the final expression of the generalized Ricci scalar curvature in the form \begin{equation}\label{RICCI} R ~~=~~ \pmatrix{ R_1 & 0 \cr 0 & R_2 \cr}~~=~~ R^{(0)} + R^{(1)} + R^{(2)}, \end{equation} where $ R^{(0)},~R^{(1)},~R^{(2)} $ represent terms proportional to $ m^0,~m,~m^2 $ respectively. The explicit expressions of $ R^{(0)}, R^{(1)}$ and $ R^{(2)}$ are given as follows: \begin{eqnarray}\label{R0} R^{(0)} &~~=~~& {1\over 2}~\pmatrix{ r_1 & 0 \cr 0 & r_2 \cr} - {1\over 32}~\Big(~ 3 \Phi^2 F^2 +2 \Phi {\tilde \Phi}{\tilde E}^{a\mu}E^\rho_{~a}{\tilde E}^{b\nu } E^\tau_{~b} {\tilde F}_{\mu \nu } F_{\rho \tau } - {\tilde \Phi}^2 {\tilde F}^2~\Big) \cr &~~~-~& {1\over 2}~{\tilde E}^{a\mu} E^\nu_{~a} {\partial_\mu \Phi \over \Phi } {\partial_\nu {\tilde \Phi} \over {\tilde \Phi}} + {1\over 2} ~{\tilde E}^{a\mu} E^\nu_{~a} {\partial_\mu {\tilde \Phi} \over {\tilde \Phi}} {\partial_\nu {\tilde \Phi} \over {\tilde \Phi}} - {1\over 2}~{\tilde G}^{\mu \nu } {\partial_\mu \partial_\nu \Phi \over \Phi} \cr &~~~-~& {1\over 2}~{\tilde E}^{\mu a}E^\nu_{~a}{\partial_\mu \partial_\nu {\tilde \Phi} \over {\tilde \Phi} } - {1\over 2}~ {\tilde E}^{a\mu}\partial_\mu E^\nu_{~a} {\partial_\nu \Phi \over \Phi} - {1\over 2}~ {\tilde E}^{a\mu} \partial_\mu E^\nu_{~a} {\partial_\nu {\tilde \Phi} \over {\tilde \Phi}} \cr &~~~+~& {1\over 4}~\Big(~{\tilde E}^\mu_{~a}{\partial_\mu \Phi \over \Phi} + {\tilde E}^\mu_{~a}{\partial_\mu {\tilde \Phi} \over {\tilde \Phi}}~\Big) \Big[~E^{a\nu} E^{b\rho}( \partial_\rho E_{b\nu} - \partial_\nu E_{b\rho}) \cr &~~~+~& {\tilde E}^{a\nu} {\tilde E}^{b\rho}( \partial_\rho {\tilde E}_{b\nu} - \partial_\nu {\tilde E}_{b\rho})~\Big], \end{eqnarray} \begin{eqnarray}\label{R1} R^{(1)} &~~=~~& m \bigg [~{1\over4}~ (\partial_\mu E_{a\nu} - \partial_\nu E_{a\mu}) \Big(~ 3 E^{a\nu} E^\mu_{~b}{\tilde E}^{b\rho}A_\rho - 4 E^{a\nu} A^\mu ({\tilde E}^{b\rho}E_{b\rho}) - 4 E^\mu_{~b}A^a {\tilde E}^{b\nu } \cr &~~~+~& 4 A^\mu E^a_{~\rho}{\tilde E}^{b\rho} E^\nu_{~b} + 8 A^\mu {\tilde E}^{a \nu } + 12 A^\nu E^{a\mu } + E^{a\mu}G^{\nu \rho}{\tilde A}_\rho - E^{b\nu} E^{a\mu}{\tilde A}_b \Phi{\tilde \Phi}^{-1} \cr &~~~-~& A^\nu E^{a\mu }{\tilde \Phi} \Phi^{-1}~\Big) +{1\over 4}~ (\partial_\mu {\tilde E}_{a\nu} - \partial_\nu {\tilde E}_{a\mu}) \Big(~4 {\tilde E}^{a\mu} A^b {\tilde E}^{\nu }_{~b} + {\tilde E}^{a\mu}{\tilde G}^{\nu \rho} A_\rho \cr &~~~-~& {\tilde A}^\nu {\tilde E}^{a\mu } \Phi {\tilde \Phi }^{-1} + {\tilde E}^{a\mu} {\tilde E}^{b\nu} E^{\rho}_{~b}{\tilde A}_\rho - {\tilde E}^{b\nu }{\tilde E}^{a\mu} A_b \Phi^{-1}{\tilde \Phi}~\Big) - {1\over 4} F_{\mu\nu}~\Big(~{5\over 2}~ {\tilde E}^{a\mu} E^\nu_{~a} \cr &~~~+~& {3\over 2}~\Phi {\tilde \Phi}^{-1} G^{\nu\rho} E^{b\mu }{\tilde E}_{b\rho } + ( {\tilde A}_\rho - A_\rho)~( 2\Phi {\tilde \Phi} G^{\mu\rho} {\tilde A}_a E^{a\nu} + 3 \Phi^2 A^\mu E^{a\nu} {\tilde E}^\rho_{~a} )~\Big) \cr &~~~+~& {1\over 4}~ {\tilde F}_{\mu \nu}~\Big(~{3\over 2} ( {\tilde A}_\rho -A_\rho) \Phi {\tilde \Phi} A^b {\tilde G}^{\mu \rho} {\tilde E}^\nu_{~b} - {1\over 2}~ {\tilde \Phi} \Phi^{-1} {\tilde G}^{\mu \rho} {\tilde E}^{a\nu }E_{a\rho} + {1\over 2} {\tilde E}^{a\mu} E^\nu_{~a} \cr &~~~+~& {\tilde \Phi}^2 ({\tilde A}_\rho - A_\rho) {\tilde A}^\mu {\tilde E}^{b\nu}E^\rho_b ~\Big) +{1\over 2}~{\tilde E}^{a\mu }\partial_\mu \Big( {\tilde E}^{\rho}_{~a} (A_\rho - {\tilde A}_\rho \Phi {\tilde \Phi}^{-1}) + E^\rho_{~a} ( {\tilde A}_\rho - A_\rho {\tilde \Phi}\Phi^{-1})~\Big) \cr &~~~-~& E^\mu_{~a} \partial_\mu \Big(~ A_\rho {\tilde E}^{a\rho} - A^a ( {\tilde E}^{b\rho} E_{b\rho}) + 3 A^a~\Big ) +{1\over 4}~ {\partial_\mu \Phi \over \Phi}~\Big( ~3A^a {\tilde E}^\mu_{~a} - A^a {\tilde E}^\mu_{~a}({\tilde E}^{b\rho}E_{b\rho}) \cr &~~~+~& 3 {\tilde E}^{a\mu}E^\rho_{~a} {\tilde A}_\rho - {\tilde A}^\mu (E^{b\rho} {\tilde E}_{b\rho}) + 3 {\tilde A}^\mu + {\tilde G}^{\mu\rho} A_\rho - 2 {\tilde E}^{a\mu} A_a {\tilde \Phi} \Phi^{-1}~\Big) \cr &~~~+~& {1\over 4}{\partial_\mu {\tilde \Phi}\over {\tilde \Phi}} {}~ \Big(~ 3A^\mu -A^\mu ({\tilde E}^{b\rho}E_{b\rho}) + G^{\mu\rho} {\tilde A}_\rho - {\tilde A}^a E^{b\rho} {\tilde E}_{b\rho}E^\mu_{~a} + 3 {\tilde A}^a E^\mu_{~a} \cr &~~~-~& {\tilde E}^{a\rho}E^\mu_{~a}A_\rho + 2\Phi {\tilde \Phi}^{-1} E^\mu_{~a} {\tilde A}^a ~\Big)~\bigg]~~ , \end{eqnarray} \begin{eqnarray}\label{R2} R^{(2)} &~~=~~& {m^2 \over 16} ~\bigg[~ \Phi^{-2}~\big(- 32 + 48 {\tilde E}^{b\rho} E_{b\rho} - 7{\tilde E}^{b\mu} E_{a\mu} {\tilde E}^{a\rho} E_{b\rho} - {\tilde G}^{\mu \nu} G_{\mu \nu } - 8 ({\tilde E}^{b\rho} E_{b\rho} )^2~\big) \cr &~~~+~& {\tilde \Phi}^{-2}~\big( E^{b\rho}{\tilde E}_{a\rho} E^{a\nu} {\tilde E}_{b\nu} - G^{\mu \nu } {\tilde G}_{\mu \nu } \big) - 2 \Phi^{-1}{\tilde \Phi}^{-1}~\big( 4 - {\tilde E}^{a\rho} E_{b \rho} {\tilde E}^b_{~\nu} E^\nu_{~a}\big)~\bigg] \cr &~~~+~& {m^2\over 4} ( {\tilde A}_\mu - A_\mu) \bigg[~ A_\rho {\tilde G}^{\mu\rho} - A_a {\tilde E}^{a\rho} E_{b\rho}{\tilde E}^{b\mu } + \Phi {\tilde \Phi}^{-1} \big( E^\nu_{~a}{\tilde E}^{a\mu} A_b{\tilde E}^b_{~\nu} - A^\mu \big)~\bigg] \cr &~~~+~& {m^2 \over 4}~\bigg[~ - 6 A_\nu {\tilde E}^{b\nu}E_{b\rho}{\tilde E}^{a \rho} A_a - A_\nu {\tilde E}^{a\nu}A_a - 2 A^2 ({\tilde E}^{b\rho}E_{b\rho})^2 + 4 A^2 ({\tilde E}^{b\rho}E_{b\rho}) \cr &~~~-~& 12 A^2 + 3 A_\nu A_a {\tilde E}^{a\nu}({\tilde E}^{b\rho}E_{b\rho}) + 3 A^2 {\tilde E}^{b\nu}E_{b\mu }{\tilde E}^{a\mu}E_{a\nu} - 3 A_a{\tilde E}^{a\rho} A_b {\tilde E}^{b\nu} G_{\rho \nu} \cr &~~~+~& 3 A^2{\tilde G}^{\mu\nu}G_{\mu\nu} + 7{\tilde A}_\rho A^\rho - 4 {\tilde A}_a A^a ({\tilde E}_{b\rho}E^{b\rho}) + 10 {\tilde A}^a A_a - {\tilde A}^\mu A_\mu ({\tilde E}_{b\rho}E^{b\rho}) \cr &~~~-~& A^\mu {\tilde A}_\mu({\tilde E}^{b\rho}E_{b\rho}) + 3 {\tilde A}^\mu A_\mu + {\tilde A}_\nu G^{\mu \nu } {\tilde A}_\mu - {\tilde A}_\mu E^{a\mu} {\tilde A}_a ({\tilde E}_{b\rho}E^{b\rho}) \cr &~~~+~& 3 {\tilde A}_\rho E^{a\rho}{\tilde A}_a +4 A_\mu{\tilde A}^\mu \Phi {\tilde \Phi}^{-1} -2 {\tilde A}^2 \Phi^2 {\tilde \Phi}^{-2} + 2 {\tilde A}_a E^{a\rho}{\tilde A}_\rho \Phi {\tilde \Phi}^{-1} \cr &~~~+~& 2 A_a {\tilde E}^{a\mu} A_\mu {\tilde\Phi} \Phi^{-1}~\bigg] - {m^2 \over 4}~\big ({\tilde A}_a\Phi {\tilde \Phi }^{-1} + A_a {\tilde \Phi} \Phi^{-1}\big)~\bigg[~ A_\nu{\tilde E}^{a\nu} - A^a ({\tilde E}^{b\rho}E_{b\rho}) \cr &~~~+~& 3 A^a + {\tilde A}_\nu E^{a\nu} - {\tilde A}^a ({\tilde E}_{b\rho}E^{b\rho}) + 3 {\tilde A}^a ~\bigg]~ \cr &~~~+~& {m^2 \over 16} \big({\tilde A}_\mu - A_\mu \big) ({\tilde A}_\nu - A_\nu)~ \bigg[~ 6\Phi {\tilde \Phi}\big( A^\mu {\tilde A}^\nu - {\tilde A}^a A_a {\tilde E}^{a\mu} E^\nu_{~a} \big) \cr &~~~+~& 5 \Phi^2 ~\big( A^2 {\tilde G}^{\mu \nu} - A_b {\tilde E}^{b\mu } {\tilde E}^{a \nu} A_a) + {\tilde \Phi}^2~\big( {\tilde A}^2 G^{\mu \nu} - {\tilde A}_a {\tilde A}_b E^{a\nu} E^{b\mu }\big)~\bigg]~~, \end{eqnarray} where \begin{equation} F_{\mu \nu }= \partial_\mu A_\nu - \partial_\nu A_\mu = \pmatrix{ f_{1\mu\nu} & 0 \cr 0 & f_{2\mu\nu} \cr} ~~, \end{equation} and $r_1$ and $ r_2$ are the ordinary Ricci scalar curvatures on the first and second copies of spacetime, respectively. The volume element is given by \begin{equation}\label{volume} D^5X~=~D^4X \sqrt{-det | {\cal G}|} \end{equation} Here $det|{\cal G}|$ denotes the determinant of our generalized metric defined in Eq.(\ref{METRIC}) and is given by \begin{eqnarray} det|{\cal G}|& \doteq &{1\over 5!}{\epsilon }_{N_1 N_2 N_3 N_4 N_5} {\epsilon}_{M_1 M_2 M_3 M_4 M_5}{\cal G}^{N_1 M_1}{\cal G}^{N_2 M_2} {\cal G}^{N_3 M_3} {\cal G}^{N_4 M_4} {\cal G}^{N_5 M_5} \cr & = & {1\over 4!}{\epsilon}_{ \nu_1 \nu_2 \nu_3 \nu_4} {\epsilon}_{\mu_1 \mu_2 \mu_3 \mu_4 } {\cal G}^{\nu_1\mu_1} {\cal G}^{\nu_2 \mu_2} {\cal G}^{\nu_3 \mu_3}{\cal G}^{\nu_4 \mu_4} {\cal G}^{55} \equiv det|G|\Phi~{\bf 1}, \end{eqnarray} where ${\epsilon}$'s are the fully antisymmetric Levi-Civita tensors and \begin{equation} det|G| ~~=~~ \pmatrix{ det |g_1| & 0 \cr 0 & det |g_2| \cr }. \end{equation} The action then is defined as \begin{eqnarray}\label{action} S &~~=~~& {1\over m. \kappa}~Tr~(\int dx^4 \sqrt{-det~G}~ R)~~,\cr &~~=~~& S_1 + S_2 ~~,\cr S_1&~~=~~& \sqrt{-det |g_1|}\varphi_1 R_1~~,\cr S_2&~~=~~& \sqrt{-det |g_2|}\varphi_2 R_2~~, \end{eqnarray} where $\kappa = 16\pi^2G^{-2}/m $ and $ G $ is the Newton constant. The integration over the discrete space follows naturally to be ${1\over m} Tr $. \setcounter{equation}{0} \section{ Mass terms:} The full action of our model (\ref{action}) contain six independent interacting fields $ e^a_{1\mu},~e^a_{2\mu},~a_{1\mu},~a_{2\mu}, {}~\varphi_1 $ and $ \varphi_2 $. Since the full expression for the Ricci scalar curvature $ R $ in Eqs.(\ref{RICCI})-(\ref{R2}) is obviously extremely complex, here we will concentrate on the massive modes in our model. We will concentrate on the gravity sector first. \subsection{ Gravity and massive tensor field} To find the mass content of the tensor field we consider the part of the action that contains only tensor fields. It turns out to be \begin{eqnarray}\label{TENSORS} R_t &~~=~~& \int dx^4~\sqrt{-det |G|}~ \bigg[ ~{1\over 2}~ \pmatrix{ r_1 & 0 \cr 0 & r_2 \cr} + {m^2 \over 16} ~\bigg(~ - 40 + 48 {\tilde E}^{b\rho} E_{b\rho} \cr &~~-~~& 7{\tilde E}^{b\mu} E_{a\mu} {\tilde E}^{a\rho} E_{b\rho} - {\tilde G}^{\mu \nu} G_{\mu \nu } - 8 ({\tilde E}^{b\rho} E_{b\rho} )^2~ \cr &~~~+~& E^{b\rho}{\tilde E}_{a\rho} E^{a\nu} {\tilde E}_{b\nu} - G^{\mu \nu } {\tilde G}_{\mu \nu } + 2 {\tilde E}^{a\rho} E_{b \rho} {\tilde E}^b_{~\nu} E^\nu_{~a}~\bigg)~\bigg]. \end{eqnarray} {}From the terms proportional to $ m^2$, we can see that $ e^\mu_{1a} $ and $ e^\mu_{2a}$ are not the fields corresponding to mass eigenstates since their products appear in these terms giving rise to mixing. To find mass eigenstates we write \begin{eqnarray}\label{EPM} E^\mu_{~a}&~~=~~& {1\over 2}~ \bigg(~e^\mu_{+a} {\bf 1} + e^\mu_{-a} r~~\bigg)~~,\cr {\tilde E}^\mu_{~a}&~~=~~& {1\over 2}~\bigg(~ e^\mu_{+a} {\bf 1} + e^\mu_{-a} r~~\bigg)~~, \end{eqnarray} and substitute for them in Eq.(\ref{TENSORS}). We note that a proper mass term has the general form $ m^2 b^{a\mu} b_{a\mu} $, where $ b^a_{~\mu}$ represents the massive tensor field. With this in mind, we find two possibilities for identifying the massive fields: i) If we choose $ e^\mu_{+a} $ as the vielbein for the metric that represents gravity, we find the mass term for the tensor field $e^a_{-\mu}$ as $ \sim 15/16 m^2 e^{a\mu}_- e_{-a\mu}$ in Eq.(\ref{TENSORS}). The terms in pure $ e^a_{+\mu} $ give a cosmological constant. In the case we are considering, these terms and the constant term cancel and consequently there is no cosmological constant. Further we note that, in the vacuum $ e_+$ is a physical field as $ e_- \rightarrow 0 $ and $ e^\mu_{+a} \rightarrow \delta^\mu_a $. ii) If we choose $ e^\mu_{-a} $ as the vielbein for the gravity metric. The same terms that give a mass to $ e^\mu_{-a}$ in the previous case, now becomes the mass terms for $ e^a_{+\mu}$. Since the terms in pure $ e^a_{+\mu}$ and the constant terms do not cancel, there is a cosmological constant in this case. In vacuum, $ e^a_{+\mu}\rightarrow 0 $ and $e^\mu_{-a} \rightarrow \delta^\mu_{~a} $. The mass term for $ e^\mu_{+a}$ in this case is $ -9/16 m^2 e^{a\mu}_+ e_{+a\mu} $. There are also quartic terms in $ e^\mu_{+a}$. It would be interesting to see whether this negative mass terms lead to spontaneous symmetry breaking patters. In the two limiting cases, when the massive tensor field is set to zero we have the usual Einstein theory with the vielbein $e^\mu_{-a}$ or the theory with the vielbein $e^\mu_{+a}$ together with a cosmological constant. Now we will consider the mass terms of the vector and scalar fields with the above two choices. \subsection{ Mass terms of vector and scalar fields} At classical level the tensor fields do not alter the mass terms of vector and scalar fields. Hence we will turn off the tensor fields and consider two limiting cases $ E^\mu_{~a} = \delta^\mu_{~a}$ and $ E^\mu_{~a}= \delta^\mu_{~a}r $. After inserting the particular $ E^\mu_{~a}$ into the the expression for $R^{(2)}$ we find: i) $ E^\mu_{~a}=\delta^\mu_{~a} $ : There is no mass terms for the scalar fields. However, the mass term for vector fields is $ 4 m^2 a_{-\mu}^2 $ where $ a_{\pm \mu} = 1/2 ( a_{1\mu}-a_{2\mu})$. This means that in this case $ a_{+\mu}$ is massless and $ a_{-\mu}$ is massive. ii) $ E^\mu_{~a} =\delta^\mu_{~a}r$ : The mass terms in this case are given by \begin{equation} R^{(2)}~~=~~ -96m^2 \varphi_{-}^{2} - 36 m^2 a_{-}^2~~, \end{equation} where $ \varphi_{\pm} = \varphi_1 \pm \varphi_2 $. The action for this part is \begin{equation} S_m ~~\sim~~ \int dx^4 -96 m^2(\varphi_{-}^{3} +\varphi_{+}\varphi_{-}^{2}) - 36 m^2( \varphi_{+} + \varphi_{-}) a_{-}^2 ~~. \end{equation} Note that in vacuum $ \varphi_+ = 1 $. Therefore $ \varphi_{-} $ is the physical mode while we have to expand $\varphi_{+}$ in terms of the physical field $\sigma $ as follows: \begin{equation} \varphi_{+} ~~=~~2 exp(-\sigma ) ~~, \end{equation} where in vacuum $\sigma \rightarrow 0 $. Using this expansion, the mass terms of $ a_{-\mu}$ and $ \varphi_{-}$ are $-36 m^2 a^2_{-}$ and $ -96m^2 \varphi^2_{-}$ respectively. These mass terms as well as the mass term for the tensor field $e^\mu_{+a}$ in this case are negative. It would be interesting to include the quartic terms to see whether these negative mass terms lead to some spontaneous symmetry breaking patterns. The quartic potential for vector fields are already there in Eq.(\ref{R2}). To have the quartic potential for the scalar field, however, one has to modify the wedge product of forms in Eq.(\ref{wedge}). Such modifications will be discussed elsewhere. \setcounter{equation}{0} \section{Summary and Conclusions:} We have in the previous papers \cite{VIWA,LVW} developed a discretized version of Kaluza-Klein theory by replacing the continuous fifth dimension by two discrete points. In the language of NCG, we may speak of two copies of spacetime instead of an infinite number of them in the standard Kaluza-Klein theory ( For every internal point in the fifth dimension we have a four-dimensional spacetime). The geometry of the extended spacetime permitted us to introduce a generalized vielbein consisting of a pair of tensor, a pair of vector and a pair of scalar fields. When we imposed the standard metric compatibility and torsion free conditions to determine the connection 1-form, we found constraints on the vielbeins in the form of dynamical dilaton fields that implied new and interesting consequences. In the present paper we have pursued the investigation further to see whether we can eliminate the constraints on the vielbeins by relaxing the torsion free condition. In order to remain as close to the Riemannian geometry as possible, we still require that the torsion 2-forms corresponding to the physical spacetime do vanish. However, by making an ansatz about torsion 2-form corresponding to the internal space, we determine uniquely not only all the connection 1-form coefficients, but also the nonvanishing torsion components in terms of the assumed vielbeins. This is in contrast to the usual Riemannian geometry where nonvanishing torsion does not lead to a unique determination of the connection coefficients. With the unique determination of the connection coefficients, we obtain a Lagrangian and an action that has a rich and complex structure with interacting tensor, vector and scalar fields. It appears as sum of two terms $ S_1$ and $ S_2$, each consisting of all the six independent fields and each representing a generally covariant action. In $ S_1 (S_2)$, the vierbein $e^\mu_{1a}~( e^\mu_{2a})$ acts as the metric field with appropriate kinetic term while the other $e^\mu_{2a}~(e^\mu_{1a}) $ coupled to $e^\mu_{1a}~(e^\mu_{2a})$ in quadratic and quartic terms. This suggests that $ e^\mu_{1a}$ and $e^\mu_{2a}$ are not eigenstates of mass. Instead we have two mass eigenstates as $e^\mu_{\pm a}= e^\mu_{1a} \pm e^\mu_{2a} $. We have two possibilities of choosing $e^\mu_{+a} $ or $ e^\mu_{-a}$ as representing the gravity field. In the first case, $e^a_{-\mu}$ and $ a_{-\mu} $ are massive fields while the scalar fields and $a_{+\mu} $ are massless. There is no cosmological constant in this case. In the second case, there is a cosmological constant and negative mass terms for tensor, vector and scalar fields. In conclusion, we like to observe that our discretized version of Kaluza-Klein theory within the framework of NCG demonstrates an extremely promising approach to internal structure of elementary particles. If the internal space is discrete, one obtains only a finite number of massive modes and thus avoids the problem of infinite number of massive modes and of the necessity of truncation. In addition to having mass, the fields have interactions proportional to the mass parameter $ m $ and the Newton constant $ G$.It is extremely interesting to explore the consequences of such theory on gravity. The highly correlated interactions also suggest strong quantum implications that are fascinating to study. \noindent {\bf Acknowledgments.} This work was supported in part by the U.S. Department of Energy under contract number DE-FG02-85ER40231. One of the authors ( N.A.V.) thanks Profs T.N.Truong and Y.X. Pham for financial support and invitation to give talks on NCG at \'Ecole Polytechnique and University Paris VI where the hospitality and stimulating discussions inspired many ideas in this paper. K.C.W. would like to thank the Fulbright Foundation for a grant and Dr.G.C.Joshi for many useful discussions and his hospitality at the University of Melbourne, Australia where this work was partially done. The contribution of G.Landi at the initial stage of this research program, mainly in Ref.\cite{LVW} is greatly appreciated. \bigskip
\section*{Introduction} \vspace*{-3.1mm} Gamma-ray bursts (GRBs) continue to confound astrophysicists nearly a quarter century after their discovery \cite{KSO73}. Before the launch of CGRO, most scientists thought that GRBs came from magnetic neutron stars residing in a thick disk (having a scale height of up to $\sim$ 2 kpc) in the Milky Way \cite{HigLin90,Hard91}. The data gathered by BATSE showed the existence of a rollover in the cumulative brightness distribution of GRBs and that the sky distribution of even faint GRBs is consistent with isotropy \cite{Meegan92,Briggs95}. This rules out a thick Galactic disk source population, with \cite{Smith94} or without \cite{MaoP92} spiral arms. Consequently, the primary impact of the BATSE results has been to intensify debate about whether the bursts are Galactic or cosmological in origin. Galactic models attribute the bursts primarily to high-velocity neutron stars in a Galactic corona, which must extend one sixth or more of the distance to Andromeda ($d_{\rm M31} \sim 690$ kpc) in order to avoid any discernible anisotropy \cite{Hak94,Hartmann94}. Cosmological models place the GRB sources at distances $d \sim 1 - 3$ Gpc, corresponding to redshifts $z\sim 0.3 - 1$. A source population at such large distances naturally produces an isotropic distribution of bursts on the sky, and the expansion of the universe or source evolution can reproduce the observed rollover in the cumulative brightness distribution \cite{Fen93}. Within the context of this workshop, we focus on Galactic corona models involving high velocity neutron stars. A recent discussion of cosmological models may be found in, e.g., Blaes \cite{Blaes94}. \section*{High Velocity Neutron Stars} \vspace*{-3.1mm} Only a few years ago scientists thought that neutron stars had velocities of 100 - 200 km s$^{-1}$ \cite{LAS82}. But recent studies show \cite{LyneLori94,FrailGW94} that as much as 50\% of neutron stars have velocities $v > 800$ km s$^{-1}$. These velocities are so high that these neutron stars escape from the Galaxy and produce a distant, previously unknown Galactic "corona." The evidence that many neutron stars have high velocities comes from two independent directions. In the first case, long-wavelength radio observations have discovered that many young radio pulsars are associated with young ($t_{\rm age} < 10^4$ yrs) supernova remnants \cite{FrailGW94}. Sometimes the young pulsar lies within the shell-like supernova remnant; sometimes it is passing through the shell, as the spectacular radio image of the "duck" supernova remnant and pulsar PSR1757-24 reveals; and sometimes the young pulsar is associated only with a comet-like "plerion," or filled remnant. In every case the pulsar lies far from the center of the remnant. These offsets imply median transverse velocities $\sim 500$ km s$^{-1}$, with $\sim$ 1/3 of the neutron stars having transverse velocities $> 1000$ km s$^{-1}$ \cite{FrailGW94}. In the second case, a new model for the electron density in the Milky Way and a greater understanding of an important observational bias that affects the determination of pulsar velocities has dramatically increased the velocities inferred for older pulsars. The new electron density model shows that the distance to, and therefore the transverse velocity of, nearby pulsars was underestimated by about a factor of two in previous models \cite{TayCordes93}. The observational bias that affects the determination of pulsar velocities arises because young radio pulsars are born close to the Galactic plane, and move rapidly away from it if their velocity is high. After some time, the pulsars that remain within detectable range are mostly those with small velocities. The strength of the bias is illustrated by the fact that the mean of the distribution of transverse velocities is $345 \pm 70$ km s$^{-1}$ for pulsars with spindown ages $\tau <$ 3 Myr, whereas it is $105 \pm 25$ km s$^{-1}$ for pulsars with $\tau \mathrel{\mathpalette\simov >}$ 70 Myr \cite{LyneLori94}. Recent studies that incorporate these discoveries yield median neutron star total velocities $\langle v \rangle_{\rm median} \sim 600$ km s$^{-1}$, with as many as half of all neutron stars having velocities $v > 800$ km s$^{-1}$ \cite{LyneLori94,Chernoff95}. These results have revolutionized our understanding of the spatial distribution of neutron stars in the Galaxy. Since the escape velocity from the Milky Way is $\approx 500$ km s$^{-1}$ in the solar neighborhood and $\approx 600$ km s$^{-1}$ in the Galactic bulge, where most neutron stars are born, all of these high velocity neutron stars will escape from the Milky Way. They form a distant, previously unknown "corona" around the Milky Way. This distant corona contains an ample population of sources which appear isotropic when viewed from the Earth. \section*{The Galactic Corona} \vspace*{-3.1mm} Prior to the launch of CGRO, many scientists believed it likely that gamma-ray bursts came from a thick Galactic disk. But, while a Galactic disk population was the most conservative and perhaps the most popular model \cite{HigLin90,Hard91}, extended halo populations have also had a long and illustrious history (see, e.g., \cite{Fish79,Jenn82,Jenn80,Shklov85}). What did exist was a consensus that gamma-ray bursts come from magnetic neutron stars in the Galaxy. There were many reasons for this, several of which we describe below. Following the discovery by BATSE that the faint bursts are distributed isotropically on the sky, Galactic halo and corona models found new flavor (see, e.g., \cite{Hartmann94,Brainerd92,DT92,LiDer:92,SmithLamb93}) as an attractive way of reconciling all of the evidence about GRBs which favors Galactic neutron stars with isotropy. However, these models were considered somewhat ad hoc, particularly by advocates of cosmological models, because no means of producing large numbers of neutron stars in an extended Galactic halo was known [see, e.g. \cite{Pac93}]. . Consequently, the debate about whether GRBs are Galactic or cosmological in origin was characterized as one between those who advocated objects which we know produce burst-like phenomena (high velocity neutron stars; see below) but which were not known to have the necessary spatial distribution (extended Galactic halo) {\it vs.} those who advocated objects which we do not know can produce burst-like phenomena (e.g., coalescing neutron star binaries or failed supernovae) but were known to have the necessary spatial distribution (cosmological). The subsequent discovery that many neutron stars have velocities high enough to escape from the Milky Way has given models a tremendous boost. Nevertheless, these models must answer several important questions: \begin{itemize} \item[\rm $\bullet$] Can a Galactic corona of high velocity neutron stars account for the isotropic sky distribution and the rollover in the brightness distribution of GRBs seen by BATSE? \item[\rm $\bullet$] Why do only high velocity neutron stars produce GRBs? \item[\rm $\bullet$] Are GRBs beamed along the direction of motion of the neutron star or, if not, why is bursting activity delayed? \item[\rm $\bullet$] Are there energy sources sufficient to power GRBs in such a model? \item[\rm $\bullet$] Can the energy needed be released over 500 Myr or more? \item[\rm $\bullet$] Can cyclotron lines formed in regions where the magnetic field is $\sim 2\times 10^{12} - 10^{13}$ G be produced by magnetic neutron stars in GRBs with luminosities $L_{\rm burst } \sim 10^{41} - 10^{43}$ erg s$^{-1}$? \end{itemize} We consider each of these questions below. \subsection*{Ingredients in High Velocity Neutron Star Models} \begin{figure}[t] \begin{tabular}{lr} {\psfig{file=mn_dot01.ps,width=5.5cm,angle=-90}} & {\psfig{file=kg_dot01.ps,width=5.5cm,angle=-90}} \end{tabular} \caption{Distribution of neutron stars with an initial kick velocity of $1000$~\hbox{km~s$^{-1}$}\ found using the Miyamoto and Nagai (1975) potential (left panel) and using the Kuijken and Gilmore (1989) potential (right panel). Note the increased concentration of stars in an extended disk due to the focusing described in the text.}\vspace{-5mm} \end{figure} \vspace*{-3.1mm} We have calculated detailed models of the spatial distribution expected for a population of high-velocity neutron stars born in the Galactic disk and moving in a Galactic potential that includes the bulge, disk, and a dark matter halo. All earlier studies of which we are aware that included these components of the gravitational potential employed the potential given by Miyamoto and Nagai \cite{MN75}. For studies of high-velocity neutron stars, it is essential to use a potential that is realistic out to very large distances. We find that the Miyamoto and Nagai \cite{MN75} implies the existence of an extended disk, far beyond the observed Galactic disk. Approximately $\approx$ 20\% of the mass lies outside $r = 20$ kpc, whereas in more realistic exponential disks less than 4\% of the mass lies outside $20$~kpc. The focusing effect of such a disk can be seen by calculating the ratio of the $z$ components of the force using the Miyamoto and Nagai \cite{MN75} potential to the force due to a point mass for $r\gg a, b, z$: \begin{equation} {F_z^{MN}\over F_z^{PM}} = 1 + {a\over(b^2 +z^2)^{1/2}}, \end{equation} where $a,b$ are parameters describing the Miyamoto and Nagai potential. Typically $a\approx 4$~kpc, $b\approx 0.2$~kpc, so that $F_z^{MN}/F_z^{PM} \rightarrow \; \approx 20$ as $z \rightarrow 0$. The use of the Miyamoto and Nagai \cite{MN75} potential distorts the orbits of neutron stars whose initial velocity vectors lie in or near the plane of the disk, and leads to an anisotropic spatial distribution that is entirely an artifact of the unrealistic disk potential (see Figure 1). We therefore use the mass distribution and potential given by Kuijken and Gilmore \cite{KG:89}. Details of our calculation are given in \cite{BulLam95a}. \begin{figure}[t] \begin{center} \begin{tabular}{lr} {\psfig{file=plo2cth.ps,width=5.5cm,angle=-90}} & {\psfig{file=plo2sb2.ps,width=5.5cm,angle=-90}} \\ {\psfig{file=plo2m31.ps,width=5.5cm,angle=-90}} & {\psfig{file=plo2pf2.ps,width=5.5cm,angle=-90}} \\ \end{tabular} \end{center} \caption{Comparison of a Galactic corona model with the inclusion of M31 in which neutron stars are born with a kick velocity of $1000$~\hbox{km~s$^{-1}$}\ and have a burst-active phase lasting $\Delta t = 500$ million years with a carefully-selected sample of 285 bursts from the BATSE 2B catalogue. Panels (a), (b), and (c) show the contours in the ($\delta t$, $d_{\rm max}$)-plane along which the Galactic dipole, Galactic quadrupole moments, and dipole towards M31 of the model differ from those of the data by $\pm$ 1$\sigma$ (solid lines), $\pm$ 2$\sigma$ (dashed line), and $\pm$ 3$\sigma$ (short-dashed line) where $\sigma$ is the model variance; the thin line in panel (a)-(c) shows the contour where the dipole moment for the model equals that for the data. Panel (d) shows the contours in the ($\delta t$, $d_{\rm max}$)-plane along which 32\%, 5\%, and $4 \times 10^{-3}$ of simulations of the cumulative distribution of 285 bursts drawn from the peak flux distribution of the model have KS deviations $D$ larger than that of the data. } \vspace{-6mm} \end{figure} \subsection*{Sky and Brightness Distributions of Bursts} \vspace*{-3.1mm} Our detailed dynamical calculations of the Galaxy show that a distant corona of high velocity neutron stars can easily account for the isotropic angular distribution and the brightness distribution of GRBs (Figure 2) [see also \cite{LiDer:92,LTD:94,Pods:94}]. \figside{\leavevmode\psfig{file=batse_pvo.ps,width=5cm,angle=-90}} {\footnotesize{\bf FIG.~\ref{batse_pvo}}. \footnotesize \baselineskip 1mm Comparison of the brightness distribution of bursts from a Galactic corona of high velocity neutron stars (thin line) and the brightness distribution of both BATSE and PVO gamma-ray bursts (thick lines) \cite{Fen93}.}{\label{batse_pvo}} In high-velocity neutron star models, the slope of the cumulative peak flux distribution for the brightest BATSE bursts and the PVO bursts reflects the space density of the relatively small fraction of burst sources in the vicinity of the Sun ($d \mathrel{\mathpalette\simov <} 50$ kpc). A spread in neutron star kick velocities, in neutron star ages at which bursting behavior begins, or in the burst luminosity function tends to produce a cumulative peak flux distribution with a slope of -3/2, the value expected for a uniform spatial distribution of sources which emit bursts that are ``standard candles.'' Figure~\ref{batse_pvo} shows that a spread of less than a factor of 10 in the luminosity function, which is consistent with what know about GRBs, is sufficient to produce agreement with not only the BATSE, but also the PVO, brightness distribution of GRBs. Beaming along the direction of motion of the neutron star can also reproduce the combined BATSE and PVO brightness distributions \cite{DLT:93,LTD:94}. The Galactic corona model predicts subtle anisotropies as a function of burst brightness, which are a signature of the model and may offer a means of verifying or rejecting it \cite{LiDer:92,DLT:93,Pods:94,BulLam95a}. \begin{figure}[th] \vspace{-3mm} \begin{center} \begin{tabular}{lr} {\psfig{file=bd.100.ps,width=4.1cm,angle=-90}} & {\psfig{file=map.100.01.ps,width=7.3cm,angle=-90}} \\ {\psfig{file=bd.300.ps,width=4.1cm,angle=-90}} & {\psfig{file=map.300.01.ps,width=7.3cm,angle=-90}} \\ {\psfig{file=bd.500.ps,width=4.1cm,angle=-90}} & {\psfig{file=map.500.01.ps,width=7.3cm,angle=-90}} \end{tabular} \end{center} \vspace{-1.5mm} \caption{Sky distribution and brightness distribution of bursts from a Galactic corona of high velocity neutron stars for BATSE sampling distances $d_{\rm max}$ = 100, 300, and 500 kpc. Note that an excess of the bursts appears only when $d_{\rm max}$~=~500~kpc.} \vspace{-7mm} \end{figure} It has often stated that Andromeda, a bright galaxy similar to our own Milky Way and lying only $700$ kpc away, imposes a severe constraint on extended halo models \cite{Hak94}. This is true, however, only if the halo extends to large distances \cite{BT87}. However, the halo of the Milky Way can extend only 1/3 - 1/2 of the distance to Andromeda because of tidal disruption. A similar statement has been thought to be true for corona models because in such models Andromeda produces its own ``wind'' of high velocity neutron stars. Some of these will travel toward us, and when they produce GRBs, BATSE should detect them. However, Andromeda imposes little constraint if the bursts are beamed along the direction of motion of the neutron star, as some models posit \cite{DLT:93,LTD:94}. Then only the rare neutron star in the corona of Andromeda whose motion is almost directly toward or away from us would be visible. So long as the BATSE sampling depth $d_{\rm max} <$ 700 kpc (the distance to Andromeda), the few bursts visible from Andromeda would always be swamped by bursts from the many high velocity neutron stars born in the Milky Way and moving away from us. Only if $d_{\rm max} >$ 700 kpc, so that a large number of the neutron stars in the Andromeda corona whose motions are away from us are visible, would an excess toward Andromeda be detectable \cite{BulCopLam:95c}. Even if the bursts radiate isotropically in all directions, detailed dynamical calculations of the motion of neutron stars in the combined gravitational potential of the Milky Way and Andromeda show that an excess of bursts toward Andromeda is not detected until one samples distances $d_{\rm max} \sim 500$ kpc from Earth (see Figure~4) \cite{Pods:94,BulCopLam:95c}. Thus there is ample parameter space (BATSE sampling distances $d_{\rm max} \approx 100 - 500$ kpc) for a population of sources in a Galactic corona. A larger sample of BATSE bursts or a more sensitive instrument might reveal an excess of bursts toward Andromeda. If so, this would constitute definitive evidence that the bursts are Galactic in origin. Lack of an excess toward Andromeda would be compelling evidence that the bursts are cosmological in origin only if made by an instrument at least 50 times more sensitive than BATSE, given the possibility that the bursts are beamed along the direction of motion of the neutron star and current constraints on the Galactic corona model. \section*{Soft Gamma-Ray Repeaters} \vspace*{-3.1mm} We have seen that a Galactic corona of high velocity neutron stars can easily account for the BATSE sky distribution and brightness distribution of gamma-ray bursts. Is there any evidence that high velocity neutron stars can produce burst-like behavior? Yes, there is. Soft gamma-ray repeaters produce high energy transients whose durations overlap with those of GRBs, and whose characteristic spectral energies form a continuum with those of GRBs. The main distinction between SGRs and GRBs is that the former have been clearly shown to repeat on time scales of days to years whereas the latter have been thought not to repeat. But recently, a number of scientists have found significant evidence that GRBs also repeat \cite{QL93,WangLing93,WangLing95,Quashnock95}. Three soft gamma-ray repeaters are known. Two lie in the Galactic disk at distances of tens of kpc (SGRs 1806-20 and 1900+14); the third lies in in the Large Magellanic Cloud in the halo of the Milky Way at a distance of 50 kpc. All three are associated with young supernova remnants \cite{Evans80,Kulkarni93,Kouvel94,Murakami94,Hurley94}. In two cases, the soft gamma-ray repeater lies far away from the center of the supernova remnant, implying a neutron star velocity of $\mathrel{\mathpalette\simov >} 1000$ km s$^{-1}$ \cite{Evans80,Hurley94} Clearly, high velocity neutron stars can produce burst-like behavior. If GRBs come from high velocity neutron stars in a distant Galactic corona, there are additional similarities between GRBs and SGRs. Both have luminosities $L \sim 10^{41}-10^{43}$ erg s$^{-1}$. Both also appear to have strong magnetic fields, as we discuss below. These similarities and the ones we discussed above suggest a physical or evolutionary relationship between SGRs and GRBs. The unification of these two phenomena is a very attractive feature of the Galactic hypothesis. \section*{The Famous 1979 March 5 Gamma-Ray Transient} \vspace*{-3.1mm} We have seen that high velocity neutron stars can produce burst-like behavior. Have high velocity neutron stars ever been seen to produce an event that looks like GRBs? The answer is ``yes.'' The event is the famous 1979 March 5 gamma-ray transient. The source of this famous event is SGR 0526-66, which lies in in the Large Magellanic Cloud in the halo of the Milky Way at a distance of 50 kpc. It is associated with the young supernova remnant N49 \cite{Evans80,Rotsch94} SGR 0526-66 lies far away from the center of the supernova remnant, implying a velocity greater than 1200 km s$^{-1}$. Seventeen bursts have been observed from this source \cite{Mazets79,Golen84}. The distribution of the durations of these bursts overlaps completely with that of GRBs. The burst had an intense spike which lasted $\sim 0.2$ s, followed by $\sim 200$ s of emission which exhibited an 8 s periodicity \cite{Mazets79}. The association with the supernova remnant N49 and the 8 s periodicity leave little doubt that this object is a neutron star. The existence of pulsations implies a strong magnetic field. The spectrum of the emission following the intense spike had a characteristic spectral energy $\langle E \rangle \approx 40$ keV, typical of SGR bursts. Although nine different satellites observed the March 5th event \cite{Evans80}, the intensity of the spike produced so-called ``dead-time'' and ``pulse pike-up'' effects which precluded reliable analyses of the spectrum. Recently, Fenimore et al. \cite{Feni95} used the power of present-day computers to unravel these effects in the ICE and PVO instruments. They found that the spike has a characteristic spectral energy $\langle E \rangle \approx 200$ keV, with no soft component, like a typical gamma-ray burst. Whether the 1979 March 5 event is a GRB or a unique event can be argued either way. But either way, it demonstrates that distant high velocity neutron stars in the Galactic halo can produce events that have the energy, the spectrum, and the duration of GRBs. This evidence strongly supports the high velocity neutron star model. \section*{Energetics} \vspace{-3.1mm} We take $F_{\rm peak} \sim 10^{-7}$ erg cm$^{-2}$ s$^{-1}$ as the typical peak energy flux of a BATSE burst. Then the typical burst luminosity is \begin{equation} L_{\rm burst} \sim 10^{41} \left({F_{\rm peak} \over {10^{-7} \; \hbox{erg} \; \hbox{cm}^{-2} \; \hbox{s}^{-1}}}\right) \left({d \over {100 \,\hbox{kpc}}}\right)^2 \; \hbox{erg} \; \hbox{s}^{-1} \; . \end{equation} Taking 5 s as the average photon flux-to-energy flux conversion factor for BATSE bursts \cite{Fen93}, the typical burst energy is \begin{equation} E_{\rm burst} \sim 5 \times 10^{41} \left({F_{\rm peak} \over {10^{-7} \; \hbox{erg} \; \hbox{cm}^{-2} \; \hbox{s}^{-1}}}\right) \left({d \over {100\, \hbox{kpc}}}\right)^2 \; \hbox{erg} \; . \end{equation} The rate of burst detection by BATSE corresponds to an all-sky rate $R^{\rm BATSE}_{\rm burst} \sim 800$ bursts yr$^{-1}$. Assuming a neutron star birth rate $R_{\rm NS} \sim 3 \times 10^{-2}$ yr$^{-1}$, each neutron star must produce a total number of bursts \begin{equation} N \approx {R^{\rm BATSE}_{\rm burst} /( {f_{\rm escape} R_{\rm NS}})} \approx 8 \times 10^4 \left({{f_{\rm escape}} / {0.3}} \right)^{-1} \; , \end{equation} during its burst-active phase, where $f_{\rm escape}$ is the fraction of neutron stars born with velocities high enough to escape from the Galaxy. Then the total supply of energy needed by each neutron star in the Galactic corona is \begin{equation} \label{etot} E \sim N ({E_{\rm burst}/ {5 \times 10^{41}}}) \sim 10^{46} \; \hbox{erg} \; , \end{equation} Among possible energy sources are gravitational energy from accretion of planetesimals \cite{Cole95,Woosley95}, crustal strain energy from spin down of the neutron star, and magnetic field energy stored in the interior of the neutron star \cite{Pods:94}. Below we discuss each possibility in turn. \subsection*{Accretion} \vspace*{-3.1mm} The gravitational energy released by accretion is \begin{equation} E_{\rm burst} = {{GM \,\Delta M_{burst}}/ R}\, . \end{equation} Taking a neutron star mass $M=1.4M_\odot$ and radius $10$~km, \begin{equation} \Delta M_{burst} \approx 10^{21} \left( {E_{\rm burst}/ {5 \times 10^{41} \; \hbox{erg}}} \right) \hbox{g} \, . \end{equation} Then the total mass needed to power the GRBs from each neutron star is \begin{equation} M \approx N \Delta M \approx 10^{27} \left({E_{\rm burst} / {5 \times 10^{41} \; \hbox{erg}}}\right) \hbox{g} \approx 10^{-6} M_\odot\, . \end{equation} This amount of mass can be supplied by a planetesimal. \subsection*{Crustal Strain Energy} \vspace*{-3.1mm} The rotational energy in a neutron star at birth is \begin{equation} E_\Omega \approx {1 \over 2} I \Omega^2 \approx 3 \times 10^{47} \left({P / {0.3\, \hbox{s}}}\right)^{-2} \hbox{erg}\, . \end{equation} However, only a fraction of this energy can be stored in the neutron star crust and released at much later time is \cite{Pods:94} \begin{equation} E_{\rm strain} \approx 0.5 \, \mu \,\theta_{\rm max}^2 \,4\, \pi \, R^2\, \Delta R_{\rm crust} \, , \end{equation} where $\theta_{\rm max}$ is the maximum strain the crust can withstand before braking, $R$ is the radius of the neutron star and $\Delta R$ is the thickness of the crust. Taking $\mu \approx 3 \times 10^{29}$ dyne cm$^{-2}$, $\theta_{\rm max} \approx 10^{-2}$, and $\Delta R_{\rm crust} \approx 0.1 R \approx 10^5$ cm, the maximum strain energy that the crust can store is \begin{equation} \label{estrain} E_{\rm strain} \approx 2 \times 10^{43} \hspace{-0.5mm}\left({\mu \over {3\times 10^{29}\, \hbox{dyne}\, \hbox{cm}^{-2}}}\right)\hspace{-0.5mm}\left({\theta_{\rm max} \over 10^{-2}}\right)^2 \hspace{-0.5mm} \left({R \over {10^6\, \hbox{cm}}}\right)^2\hspace{-0.5mm}\left({{\Delta R_{\rm crust}} \over {0.1 R}}\right) \hbox{erg} \; . \end{equation} This energy is much smaller that given by equation \ref{etot}. Thus the strain energy that can be stored in the neutron star crust as it solidifies while the neutron star is rotating rapidly appears unable to supply the total energy needed to power the bursts in the Galactic corona model. \subsection*{Magnetic Field Energy} \vspace*{-3.1mm} We know from accretion-powered pulsars and rotation-powered pulsars that the surface fields of most neutron stars lie in the range $B_s \sim 10^{11} - 10^{13}$ G. We have virtually no knowledge about the internal magnetic fields of neutron stars. If the internal field exceeds $10^{16}$ G, then superconductivity is quenched and the total energy stored in the internal magnetic field is \begin{equation} E^{\rm normal}_{\rm magnetic} \approx {{4\pi} \over 3} R^3 {B^2 \over {8\pi}} \approx {1 \over 6} R^3 B^2 \approx 10^{49} \left({B \over {10^{16}\, \hbox{G}}}\right)^2 \hbox{erg} \; . \end{equation} If the internal magnetic field is less than $10^{16}$ G, it is expected that the interior of the neutron star will be superconducting. Then the total energy stored in the internal magnetic field is \cite{Pods:94} \begin{equation} E^{\rm super}_{\rm magnetic} \approx {1 \over 6} R^3 B B_c \approx 10^{49} \left({B \over {6 \times 10^{13}\, \hbox{G}}}\right)\left({B_c \over {10^{16} \, G}}\right) \hbox{erg} \; . \end{equation} However, the energy stored in the interior magnetic field might then be released on the time scale, \begin{equation} \tau_{\rm spindown} \approx 10^7 P^2 (B/10^{12}\, \hbox{G})^{-2}\, \hbox{yr} \ll 5 \times 10^8 \, \hbox{yr} \; , \end{equation} if the spin vortices in the superfluid drag the magnetic flux tubes toward the surface of the neutron star\cite{Srinivasan90}. Thus, if the interior of the neutron star is superconducting, the amount of energy stored in the interior magnetic field is sufficient to power the bursts in the Galactic corona model but may be released over a period of time much less than the required lifetime of such sources. If magnetic field instabilities stress the neutron star crust, then from equation \ref{estrain} above, the energy released would be \cite{Pods:94} \begin{equation} E_{\rm strain} \approx 2 \times 10^{42} \left({\mu \over {3 \times 10^{29} \,\hbox{dyne}\, \hbox{cm}^{-2}}}\right) \left({\theta \over 10^{-3}}\right)^2 \left({R \over {10^6 \, \hbox{cm}}}\right)^2 \left({{\Delta R_{\rm crust}} \over {0.1 R}}\right) \hbox{erg} \; , \end{equation} which is about right for GRBs. \section*{Radiative Processes} \vspace{-3.1mm} \subsection*{Pair Fireballs} \begin{figure}[t] \begin{center} \begin{tabular}{lp{4cm}} \raisebox{-4.3cm} {\leavevmode\psfig{file=windtb.ps,width=7cm,angle=-90}} & {\footnotesize{\bf FIG.~\ref{windtb.ps}} Neutron star magnetosphere, showing the region at the magnetic pole where super-Eddington luminosities may drive a relativistic pair wind, and the region at the magnetic equator where super-Eddington luminosities may create a trapped pair fireball \cite{TD95}.} \end{tabular} \end{center} \refstepcounter{figure} \label{windtb.ps} \vspace{-8mm} \end{figure} \vspace*{-3.1mm} Our discussion of pair fireballs follows that of M\'esz\'aros \cite{Mesz95}. A compactness parameter $ \tau_{\gamma \gamma} \approx ({{L_{burst} \sigma_T} / {4 \pi r c \epsilon_\gamma \Gamma^2}})\mathrel{\mathpalette\simov <} 1$ is required for photons to be observed above an energy $\epsilon_\gamma$, where $r$ is the radius of the source and $\Gamma$ measures the relativistic velocity. For $r \approx R$, $\tau_{\gamma \gamma} \gg 1$ at $\epsilon_\gamma \approx 1$ MeV unless $\Gamma \mathrel{\mathpalette\simov >} 10^4$. This result suggests that the initial stages of pair fireball in a gamma-ray burst is optically thick. If so, we expect that the sudden release of the energy that powers the burst will produce a trapped fireball in the region of closed magnetic fields lines in the neutron star magnetosphere and a mildly relativistic wind from the regions of open field lines at the magnetic poles (see Figure 5). This picture is similar in many respects to the soft gamma-ray repeater model of Thompson and Duncan \cite{TD95}. The energy required in a solid angle $\theta$ is \begin{equation} E_{\rm burst} \approx 10^{41}\left({F_{\rm peak} \over {10^{-7} \,\hbox{erg}\, \hbox{cm}^{-2}\, \hbox{s}^{-1}}}\right) \left({d \over {100 \,\hbox{kpc}}}\right) \theta^2 \, \hbox{erg} \; . \end{equation} Pair production occurs if $\epsilon_\gamma > 4 (m_e c^2)^2 \epsilon_t^{-1} \alpha^{-2} \; , $ where $\epsilon_t$ is the lab frame target photon energy and $\alpha$ is the relative angle between the two photons. Causality implies $\alpha \mathrel{\mathpalette\simov <} \Gamma^{-1}$, or $ \Gamma^{-1} \mathrel{\mathpalette\simov <} \alpha \mathrel{\mathpalette\simov <} 2 m_e c^2 (\epsilon_t \epsilon_\gamma)^{-1/2} \; , $ and \begin{equation} \epsilon_\gamma \mathrel{\mathpalette\simov <} 10^4 (\epsilon_t/\hbox{MeV})^{-1} (\Gamma/10^2)^2 \hbox{MeV} \; . \end{equation} This implies relativistic expansion and therefore beaming. How large might $\Gamma$ be? In AGN, an initial value (near the central source) as large as $\Gamma\sim 10^4$ is often assumed. If the wind is powered by Compton scattering, a value $\Gamma \approx 10$ or so is expected. If $(E_{\rm burst}/\delta m_{\rm burst} c^2 )\mathrel{\mathpalette\simov <} 1$, where $\delta M_{\rm burst}$ is the amount of baryonic matter entrained in the outflow, then the wind will be subrelativistic due to baryonic poisoning. \subsection*{Cyclotron Lines} \vspace*{-3.1mm} Almost fifteen years ago Mazets et al. \cite{Mazets81,Mazets82} reported seeing single lines in the spectra of GRBs at low energies ($E \mathrel{\mathpalette\simov <} 70$ keV). Later Hueter \cite{Hueter88} reported single lines at low energies in the spectra of two bursts seen by HEAO-1 A4. However, the statistical significance of the lines was modest. More recently, equally-spaced lines were seen by {\it Ginga} in the spectra of three bursts \cite{Murakami88,Fenimore88,Graziani92,Yoshida92} with high significance \cite{Fenimore88,Graziani92,Freeman95}. The line features in these three bursts have been studied extensively, and there is no doubt that they exist. Lines have not been definitively seen by BATSE \cite{Palmer94}, but this fact does not strongly contradict earlier observations \cite{Band94}. Similar line features are seen in the spectra of accretion-powered pulsars\cite{MakiMih92}, which are known to be magnetic neutron stars. The equally-spaced lines seen in GRBs and in accretion-powered pulsars are easily explained in terms of cyclotron resonant scattering in a strong magnetic field \cite{Wang89,B92}. Magnetic neutron stars in the Galactic corona appear able to produce cyclotron lines even though the luminosities of the bursts might greatly exceed the so-called Eddington luminosity at which radiation pressure and gravity balance. Cyclotron lines may form, for example, in a relativistic wind flowing out from the magnetic poles of the neutron star \cite{Miller91}, or at the magnetic equator \cite{Freeman95} where hot plasma is trapped by the magnetic field \cite{Lamb82,Katz82,Katz95,TD95}. \section*{Conclusions} \vspace{-3.1mm} Detailed dynamical calculations show that a distant Galactic corona of high velocity neutron stars can easily account for the isotropic angular distribution and the brightness distribution of GRBs. Gravitational potential energy from accretion and magnetic field energy seem the most promising sources of energy which are capable of powering bursts from such a population of neutron stars. In a GRB hot plasma will likely flow from the polar cap in a relativistic wind, but will be trapped at the magnetic equator by the magnetic field. Cyclotron lines might form in either region. \subsection*{Future Prospects} \vspace*{-3.1mm} Below we mention several key observations that might confirm or refute the hypothesis that the GRBs come from a distant Galactic corona of high velocity neutron stars. {\bf Sky distribution.} Our ability to detect or place upper limits on any anisotropies in the burst sky distribution, especially as a function of burst brightness, will increase slowly but steadily as BATSE detects more bursts. Confirmation of significant Galactic dipole and/or quadrupole moments as a function of burst brightness, or overall, would provide definitive evidence that the bursts are Galactic. Further limits on any angular anisotropy will constrain, and might rule out, the Galactic hypothesis. However, the limits that BATSE will be able to achieve are not likely to be definitive, since the angular distribution of bursts from the distant Galactic corona can be very isotropic. Detection of a concentration of bursts toward Andromeda, either by BATSE, or by a more sensitive experiment would constitute definitive evidence that the bursts are Galactic in origin. Lack of an excess toward Andromeda would be compelling evidence that the bursts are cosmological in origin only if made by an instrument at least 50 times more sensitive than BATSE, given the possibility that the bursts are beamed along the direction of motion of the neutron star and current constraints on the Galactic corona model {\bf Cyclotron lines.} Other spectroscopy instruments are now operating (TGRS and Konus on Wind) or will soon be flown (e.g., HETE, Konus on Spectrum X-Gamma, etc.) which will search for lines. Further confirmation of the existence of cyclotron lines would provide strong evidence in favor of the Galactic hypothesis. {\bf Repeating}. The new bursts in the third BATSE catalog are not expected to suffer from the same limitations which afflicted bursts in the second year of observations due to failure of the tape recorders on board the {\it Compton Gamma-Ray Observatory}. It is therefore expected that the third BATSE catalogue will provide an excellent opportunity to test the repeating hypothesis. Confirmation of repeating would doom most cosmological models. \vspace{-3.1mm}
\section{Introduction} The polarimetry of the interstellar medium (hereafter ISM) testifies that dust grains are aligned. The original explanation of this phenomenon in Davis \& Greenstein \shortcite{dg} as being due to paramagnetic relaxation was later criticized in Jones \& Spitzer \shortcite{js} as inadequate unless enhanced imaginary part of grain magnetic susceptibility was assumed. At the same time, alignment involving purely mechanical processes, pioneered in Gold (1951, 1952), was shown to have its own problems (see Davis 1955, Purcell 1969, Purcell \& Spitzer 1971). A new important contribution to the field was done by Roberge \& Hanany (1990), who proposed an interesting idea, that grain alignment can be caused by ambipolar diffusion. A comprehensive study of alignment of thermally rotating grains through the ambipolar diffusion is given in Roberge, Messinger, \& Hanany (1995). Such an alignment is likely to be important for molecular clouds. However it is possible to show, that supersonic ambipolar diffusion cannot persist over extensive regions of the ISM. A really profound step in elucidating the dynamics of the ISM dust grains was done by Purcell (1975, 1979) who introduced a concept of suprathermal rotation. However, it was shown in Spitzer \& McGlynn \shortcite{sm} that this only marginally improves the efficiency of paramagnetic alignment for high rates of resurfacing. In fact, specialists in the field believe that `a relaxation mechanism orders of magnitude more efficient than that given by normal paramagnetism' is required \cite{chg}. This paper was preceded by our earlier papers devoted to the same issue. First, in Lazarian (1994) (henceforth Paper I) we provided a quantitative description of mechanical alignment of non-spherical grains. However, at that point it was unclear whether grains can be aligned mechanically, if they rotate at suprathermal velocities. Therefore, only alignment of thermally rotating grains was discussed. Then, in Lazarian (1995a) (henceforth Paper II) we addressed the problem of paramagnetic alignment of suprathermally rotating grains and showed that diffusion of oxygen over grain surface and existence of potential barrier for H$_{2}$ formation (Tielens \& Allamandola 1987) bring into being the grain critical size $l_{cr}$ and the critical number of active sites $\nu_{cr}$. It was found out that for grains with sizes less than $l_{cr}$ suprathermal rotation due to H$_{2}$ formation is suppressed and that the poisoning of active sites is accelerated when the number of active sites becomes greater than $\nu_{cr}$. However, only the case $\nu<\nu_{cr}$ was quantitatively discussed and no processes of desorption were invoked; this is a disadvantage of the latter study. The aim of this paper is to determine roughly the relative importance of processes, that can provide alignment of suprathermally rotating grains. Therefore the contribution is rather miscellaneous. First of all, in Sect.~2 we remind our reader of the Purcell's concept of suprathermal rotation. Then, we address in Sect.~3 mechanical processes, which are usually believed to be incapable of aligning suprathermally rotating grains. We show that this is not true and discuss a process that can provide efficient alignment of such grains. In Sect.~4 we address the paramagnetic alignment and discuss poisoning of grain active sites when $\nu>\nu_{cr}$ and the processes of desorption of molecules blocking the access to active sites. To provide a more comprehensive picture of alignment of suprathermally rotating grains, we touch upon the alignment of helical grains, grains subjected to radiative flows, as well as alignment due to grain magnetic moments. \section{Suprathermal rotation of grains} Here we discuss the suprathermal rotation due to formation of H$_{2}$ molecules on the ISM grains. The number of molecules ejected per second from an individual site is $\sim \gamma_{1} l^{2}n_{H}v_{1}\nu^{-1}$, where $\gamma_{1}$ is the portion of H atoms absorbed by a grain of diameter $l$ while $\nu$ is the number of sites. A recoil from a molecule departing the grain is $m_{H2}v_{H2}= (2 m_{H2} E_{H2})^{1/2}$, where $E_{H2}$ is the kinetic energy of H$_{2}$ molecule ($0.2$~eV) \cite{williams1}. Then the average squared $z$ component of torque is \begin{equation} \langle[M_{z}]^{2}\rangle\approx\frac{\gamma_{1}^{2}}{32}l^{6}n^{2}_{H}m_{H2} v_{1}^{2}E_{H2}\nu^{-1}. \label{7} \end{equation} Thus an individual grain spins up to high angular velocities, limited only by friction forces. As a result, the attainable value of angular velocity $\Omega$ is proportional to $\langle[M_{z}]^{2}\rangle^{1/2}\kappa^{-1}$, where $\kappa$ is the constant of rotational friction $\sim I_{z}t_{d}^{-1}$. Note, that $I_{z}$ is the $z$ component of the momentum of inertia and $t_{d}$ is the rotational dumping time. The latter is \begin{equation} t_{d}\approx C t_{m}= C\frac{\varrho_{s}V}{S\Phi}, \label{e.4} \end{equation} where $t_{m}$ is the time that takes a grain to collide with gaseous atoms of the net mass equal to the mass of the grain, $V$ and $\varrho_{s}$ are grain volume and density respectively; $S$ is the grain cross-section for a supersonic flux, the coefficient $C=0.6$ is a precise result for a spherical grain (Purcell \& Spitzer 1971) and $\Phi$ is the mass flux which is $v_{1}nm_{a}$ for supersonic $v_{1}$. Therefore assuming $\gamma_{1}=0.2$ and $n_{H}=n$, one obtains $\Omega\approx 2\cdot 10^{8}$~s$^{-1}$, for $\nu=10^{2}$. Some grains are not subjected to the torques caused by H$_{2}$ formation. For instance, it was shown in Paper II that those are aromatic carbonaceous grains, as H$_{2}$ molecules are being formed is states of low excitation over their surface (Duley \& Williams 1993). Nevertheless, such grains can rotate suprathermally, for instance, due to variations of the accommodation and/or photoelectric emission coefficients (see Purcell 1979). Further on, we will show that unlike suprathermal rotation due to H$_{2}$ formation, some of those processes may result in alignment on their own. Our estimates of $\Omega$ above are relevant to the long-lived spin-up (Spitzer \& McGlynn 1979). For short-lived spin-up, namely, the correlation time of Purcell rockets $t_{L}$ is much less than $t_{d}$ our estimate of $\Omega$ must be multiplied by $\sqrt{t_{L}/t_{d}}$. Speaking about Purcell alignment we will be mostly concerned with the long-lived spin-up. Paramagnetic alignment corresponding to the short-lived spin-up does not differ much from the Davis-Greenstein one corresponding to the enhanced temperature of grain rotation $KT_{gas}$, where \begin{equation} K=\frac{I_{z}\langle[M_{z}]^{2}\rangle t_{d}t_{L}}{kT_{gas}}. \label{3} \end{equation} It was pointed out by Bruce Draine (private communication), that such an enhancement can be essential in molecular clouds where temperatures of gas and grains are very close. In fact, suprathermal rotation can arise not only from H$_{2}$ formation, but also by the variations of the accomodation coefficient. We discuss these such processes in Section~5, as we show, that such variations can cause mechanical alignment of its own irrespectively from the action of the paramagnetic relaxation. \section{Mechanical alignment} \subsection{Supersonic motions} To produce mechanical alignment grains should drift supersonically in respect to gas. In Paper I it is showed that Alfv\'{e}nic perturbations are efficient in providing supersonic drift perpendicular to magnetic field lines. Two main phenomena may be invoked: Alfv\'{e}nic waves and ambipolar diffusion. It is generally accepted, that supersonic drift by ambipolar diffusion is present within sufficiently strong shocks (see fig.1 in Pilipp et al 1990). However, this cannot be the cause of alignment for the majority of the ISM grains. Indeed, it was pointed out by B.Draine, that the dissipation due to ion-neutral streaming with velocity $v_{in}$ is approximately \begin{equation} G=0.5n n_{i} \langle \sigma_{T}v_{T}\rangle m_{a} v_{in}^{2} \end{equation} which for supersonic $v_{in}=1$km s$^{-1}$ provides a dissipation rate per H atom $G/n \approx 1.7 \times 10^{-25}$erg s$^{-1}$, which is well in excess of the energy that supernova is likely to eject into the turbulent motions. In other words, the ambipolar diffusion may be important for molecular clouds and localized regions in within diffuse clouds, but for the large scale alignment other processes must be invoked. It was hinted in Paper I that large grains are inertial and therefore should decouple from ionized gas in the course of high frequency Alfv\'{e}nic oscillations. Although this process of alignment does not entail strong dissipation, it is possible to show that the limitations imposed on grain size by both damping of high frequency Alfv\'{e}nic waves and grain charge are rather strict. Indeed, if a grain without electric charge is placed within partially ionized gas subjected to Alfv\'{e}nic perturbations, e.g. with velocity $v=v_{0}\sin\omega_{A}t$, it is easy to see that the amplitude of grain drift scales as $v_{0}/\sqrt{1+\mu^{2}}$, where $\mu \approx 0.4 (\rho_{n}/\rho_{g})(v_{0}/(l\omega_{A}))$ and $\rho_{n}$ and $\rho_{g}$ are, respectively, gas and grain densities. Therefore $\mu < 1$ provides the first constrain. In its turn, Alfv\'{e}nic frequency cannot be arbitrary, as the waves of frequency higher than that of collisional frequency for ions are critically damped (see McKee et al 1994). The corresponding value of $\omega_{cr}$ provides the minimal value of $\mu$. Another constrain stems from the grain charge, as if $\omega_{A}$ is less than the grain Larmour frequency, the drift velocity scales as $v_{0}(\omega_{A}/\omega_{L})$. For a given grain size only oscillations within the range $\omega_{L}<\omega_{A}<\omega_{cr}$ are efficient in the sense of providing supersonic drift. Therefore, even if the overall non-thermal line broadening of emission lines is observed, it is not clear a priori for a given grain, that this inequality is fulfilled and that the Alfv\'{e}nic motions within aforementioned range of frequencies are supersonic. The estimates in Lazarian (1995b) show, that for the ionization ratio $10^{-4}$ it is difficult to expect grains with radii less than $10^{-5}$ cm to experience supersonic drift in the diffuse ISM, if grains are collisionally charged (see Spitzer 1978, Draine \& Sutin 1987). It is gratifying, that such a dichotomy is observed and grains of sizes less than $10^{-5}$ cm are not aligned (Kim \& Martin 1994, 1995). Radiation pressure can also cause grain drift. For this drift to be supersonic the following inequality $Q_{\it ext}P_{\it rad}>4P_{\it gas}$ should be satisfied (Purcell 1969), where $Q_{\it ext}$ is the ratio of the optical to geometrical cross sections, while $P_{rad}$ and $P_{\it gas}$ are, respectively, radiation and gaseous pressures. For hydrogen with density 10~cm$^{-3}$ and temperature 100~K, $P_{\it gas}\approx 10^{-13}$~dyn/cm$^{-2}$. Usually $Q_{\it ext}<1$ and this means that it is difficult to obtain sufficient radiation pressure to account for large-scale ISM polarization. Thus supersonic drift due to radiation pressure is confined to regions in the vicinity of bright sources. As grains carry charge, their motion under the radiation pressure is usually constrained to following magnetic field lines (Spitzer 1978, p.~202). It is shown in Habing {\it et al}. (1994) that grains should move with supersonic relative velocities within the outflows around cool giants (carbon-stars, Mira variables and OH/IR stars) due to radiation pressure. The velocities of metallic grains should be greater than those of dielectric grains in such circumstances (Il'in 1994) and this may have observational consequences in view of mechanical alignment. To summarize, supersonic drift is an essential feature of interstellar grain dynamics. In the vicinity of stars this drift takes place mostly due to radiation pressure and happens along magnetic field lines. Within molecular and atomic clouds at sufficient distances from bright sources, this drift takes place due to Alfv\'{e}nic perturbations and grains move mostly perpendicular to magnetic field lines. \subsection{Cross section alignment} Grain alignment can be quantitatively described by the following measure: \begin{equation} \sigma_{J}=\frac{3}{2}\left\langle\cos^2\beta-\frac{1}{3}\right\rangle, \label{star} \end{equation} where $\beta$ is the angle between the angular momentum and the alignment axis. The latter can coincide with the direction of a corpuscular flux, if precession on the time scale of alignment is negligible, or with the direction of magnetic field, if the period of grain precession is much smaller than the said time scale. For suprathermally rotating grains the axis of major inertia is tightly coupled to the angular momentum (Purcell 1979). Therefore for oblate spheroidal grains, $\sigma_{J}$ coincides with the Rayleigh reduction factor $\sigma$ introduced in Greenberg (1968). For prolate spheroids it is easy to show that $\sigma=-0.5\sigma_{J}$. One of the major questions that we need to answer here is whether suprathermally rotating grains can be aligned by a supersonic flow. A naive answer would be ''no`` as such rapidly rotating grains resemble gyroscopes and therefore should not be sensitive to gaseous bombardment. However, this answer ignores the role of crossovers. Further on we will show, that when both random and regular torques are dominated by recoils caused by H$_{2}$ formation, an efficient alignment of non-spherical grains is possible. We remind our reader, that during crossovers grain angular velocity becomes close to zero and a substantial randomization of $\bf J$ is present. A comprehensive theory of crossovers is given in Spitzer \& McGlynn (1979) and further on we will use the results of this study. Let us assume, for simplicity, that randomization is complete during a crossover, which is apparently true for sufficiently small grains (see Paper II for an explicit expression of the disorientation parameter). An incomplete randomization only alters the time of alignment. For instance, grain may need not one, but several crossovers to come to the state uncorrelated with an initial one. This, however, does not change anything for our further treatment. If $t_{L}$ is a characteristic correlation time for the existence of the sites of H$_{2}$ formation, the mean time back to crossover is given by \begin{equation} t_{x}\approx 1.3(t_{d}+t_{L}) \label{5} \end{equation} (Purcell 1979). Usually, it is assumed that the time of the existence of active sites is the time necessary for accreting a monolayer of refractory material (Spitzer \& McGlynn 1979). Another process that can determine $t_{L}$ is poisoning of the active sites (see Paper II and Sect.~4 of this paper). However, for our simplified treatment we do not distinguish between the two processes limiting the life time of Purcell's rockets. Indeed, what is important for us is that, for a chosen grain, $t_{L}$ is the time of accreting of $N$ heavy atoms. The latter value depends on particular processes involved, but we will see that this is not critical for the mechanism below. Consider at first a toy model, namely, assume that the axis of suprathermally rotating oblate grain can have only two positions, namely perpendicular ($\bot$) and parallel ($\parallel$) to the axis of a gaseous flux. In this model, grain axis stays for the time $t_{x}$ in one position and then undergoes the crossover and has equal chances to get either the same or the perpendicular alignment (disorientation is complete!). In the absence of the gaseous flux time scales $t_{x \parallel}$ and $t_{x \bot}$ are the same and, naturally, there is no preferential position. However, if the gaseous supersonic flux is present, $t_{d}$ given by Eq.~(\ref{e.4}) is inversely proportional to gas-grain cross sections $S_{n \parallel}$ and $S_{n \bot}$ in the two positions. Moreover, it is natural to assume that the number of heavy atoms adsorbed by the grain is proportional to the overall number of atoms striking it. Thus $t_{L}$ and $t_{x}$ are both inversely proportional to the cross section. Assuming that the time of the crossover is negligible as compared with the time of a spin-up, we conclude that for an individual grain the time averaged probability of finding the grain in a position $\bot$ or $\parallel$ is inversely proportional, respectively, to $S_{n \bot}$ and $S_{n \parallel}$. To find the constant of proportionality one needs to recall, that the probability of finding the grain in either of two positions is unity. According to the ergodic hypothesis this probability coincides with the ensemble averaged one. The above considerations can be generalized. It is easy to see, that for a continues distribution of axis positions the probability of finding grain axis within at a particular angle is inversely proportional to the cross section corresponding to this angle. It was assumed above, that both $t_{L}$ and $t_{d}$ are controlled by the same process, namely, by the interaction of a gaseous flow with a grain. One may imagine situations when $t_{L}$ and $t_{d}$ are controlled by {\it different} physical processes. If one of the times $t_{L}$ or $t_{d}$ is much less than the other, \footnote{We should bear in mind that Eq.~(\ref{5}) is an approximate one and it is checked in the range of $0.1<t_{L}/t_{d}<10$ (Purcell 1979).} the longest of the two controls the alignment. For instance, for $t_{L}\ll t_{d}$ we deal with alignment by friction, which may remind one the the idea suggested in Salpeter \& Wickramasinghe (1969). However, these processes are different, as we deal here with suprathermally rotating grains, while Salpeter \& Wickramasinghe discussed alignment of grains with enhanced rotational temperature. It is possible to show, that the alignment in the latter case is rather marginal. In the opposite limiting case, namely, $t_{L}\gg t_{d}$ we show in Sect.~5 that alignment due to photodesorption may be efficient. We do not dwell upon all these interesting possibilities here, as believe, that the alignment of suprathermally rotating grains under the simultaneous action of several processes deserves a separate study. Consider the alignment of grains subjected to Alfv\'{e}nic perturbations. If, for the sake of simplicity, we approximate grains by thin discs the cross section will vary as \begin{equation} S_{n}=\pi r^{2} |\sin\varphi \cos\psi| \end{equation} where $\varphi$ is the angle between the disc axis and magnetic field, $\psi$ is an angle in the plane perpendicular to the magnetic field and $|..|$ denote that we take the absolute value of the trigonometric functions. To account for the Larmour precession, we should perform averaging over $\psi$. Therefore, \begin{equation} \langle \cos^{2}\varphi \rangle=C \int_{0}^{2\pi}\frac{d\psi}{|\cos\psi|} \int_{0}^{\pi}\cos^{2}\varphi d\varphi \end{equation} where the normalization constant $C$ is \begin{equation} C^{-1}=\pi \int_{0}^{2\pi}\frac{d\psi}{|\cos\psi|}. \end{equation} In short, it is easy to see that $\langle \cos^{2}\varphi \rangle=0.5$ and the Rayleigh reduction factor (Greenberg, 1968), is equal to $0.25$. Grain long axis tends to be perpendicular to magnetic field, but the alignment is not perfect (compare Paper I). In another important case, when flakes stream along magnetic field lines, similar computations provide $\sigma=-0.5$, which corresponds to the perfect alignment with grain long axis along magnetic field. More efficient alignment for streaming along magnetic field as compared with Alfv\'{e}nic perturbations is a consequence of the fact, that in the former case the Larmour precession does not change the grain -- gas cross section. It easy to see, that this type of alignment is efficient for flakes, as the cross section difference averaged over the period of rotation is maximal for such grains. Contrary to this, the difference in cross sections is not large for prolate grains. For instance, for needles the ratio of the maximal to minimal averaged cross sections is just $\pi/2$ and the $\sigma$ for the most favorable conditions (a flow parallel to the magnetic field lines) is about $0.07$. Therefore the alignment of prolate grains is marginal due to the mechanism. It was implicitly assumed above that the angular momentum associated with gaseous bombardment is not important during a crossover. It is possible to show that, this is true unless grain drift velocity is comparable with velocity of H$_{2}$ molecules and/or accommodation coefficient is substantially different from unity and/or atomic hydrogen is largely converted into molecular form. A detailed discussion of the effects of gaseous bombardment is given in Lazarian (1995c). \section{Purcell alignment} We will call paramagnetic alignment of suprathermally rotating grains the Purcell alignment to distinguish it from the Davis-Greenstein mechanism acting on thermally rotating grains. Various aspects of the Purcell alignment were addressed in Paper II. Here we extend one aspect of the aforementioned analysis, namely, the one dealing with poisoning of active sites. Back in Paper II it was established that a critical number of active sites exists. This number, $\nu_{cr}$, is the mean number of sites with chemically adsorbed H atoms, that a hydrogen atom arriving to the grain surface can visit before reacting at any of these sites. It is easy to see that, if the number of active sites is less than $\nu_{cr}$, there cannot be on average more than one active site, which is left empty since a recent H$_{2}$ formation. However, if the number of active sites $\nu$ is greater than $\nu_{cr}$, the number of empty sites scales as $\nu/\nu_{cr}$ for $\nu_{cr}\gg 1$. These empty sites are the primary targets of oxygen atoms hopping over grain surface and it was observed in Paper II that for $\nu>\nu_{cr}$ poisoning increases. Therefore only the case of $\nu<\nu_{cr}$ was discussed. Here we study the case of $\nu>\nu_{cr}$. Similarly to Paper II we assume that oxygen is being immobilized on hydrogenation (Leitch-Devlin \& Williams 1984, Williams, private communication). If $\nu/\nu_{cr}$ is the expected number of empty active sites, the probability of an oxygen atom to fill any of them as a result of an individual hop is $\nu/(\nu_{cr}N_{ph})$, where $N_{ph}$ is the number of sites of physical adsorption. Therefore the probability of an empty site not to be filled in one hop of an oxygen atom is $(1-\nu/(\nu_{cr}N_{ph}))$; the same probability for $m_{h}$ hops is \begin{equation} \left(1-\frac{\nu}{\nu_{cr}N_{ph}}\right)^{m_{h}} \approx \exp \left(-\frac{\nu}{\nu_{cr}} \frac{m_{h}}{N_{ph}}\right). \label{ff} \end{equation} Therefore the characteristic time of poisoning of $\nu/2$ active sites is \footnote{ We assume that not more than one oxygen atom hop over grain surface at any particular time. This may not be true for dense cores of molecular clouds, where the concentration of atomic hydrogen is low.} \begin{equation} t_{p}=\nu t_{O}\frac{1}{1-\exp \left(-\frac{\nu}{\nu_{cr}} \frac{m_{h}}{N_{ph}}\right)}, \end{equation} where $t_{O}$ is the time of an oxygen atom arrival at grain surface. For $\frac{\nu}{\nu_{cr}}\frac{m_{h}}{N_{ph}}\ll 1$ it is possible to expand the exponent in Eq~(\ref{ff}) \begin{equation} t_{p} \approx \nu_{cr} t_{O} \frac{N_{ph}}{m_{h}}\left[1+\frac{1}{2}\frac{m_{h}} {N_{ph}} \frac{\nu}{\nu_{cr}}-O\left(\frac{m_{h}^{2}}{N_{ph}^{2}} \frac{\nu^{2}} {\nu_{cr}^{2}}\right)\right]. \end{equation} The number of hops $m_{h}$ is the ratio of the time required to hydrogenate oxygen, which we denote $t_{ch}$ to the time of an individual hop $t_{h}\approx 10^{-12}\exp(E_{h}/(kT_{s}))$ s, where $E_{h}$ is the energy of potential barrier and $T_{s}$ is the grain temperature. When the number of active sites exceeds $\nu_{cr}$ any hydrogen atom scans only the fraction $\nu_{cr}/\nu$ of the entire surface before either reacting with another H atom at the active site or being trapped by an empty active site. The probability, that O atom adsorbed by the grain is present over this part of the surface is proportional to $\nu_{cr}/\nu$. Therefore the time hydrogenation of oxygen is of the order of the $\nu/\nu_{cr}$ over the timescale of hydrogen arrival. Therefore \begin{equation} t_{p}\approx \nu_{cr}\frac{\nu_{cr}}{\nu}\frac{t_{O}}{t_{H}}t_{h}N_{ph} \left[1+\frac{1}{2}\frac{\nu^{2}t_{H}}{\nu_{cr}^{2}t_{h}N_{ph}}- O\left(\frac{\nu^{4} t_{H}^{2}}{\nu_{cr}^{4}t_{h}^{2}N_{ph}^{2}}\right)\right], \label{12} \end{equation} where the ratio $t_{O}/t_{H}=(\gamma_{1}n_{H}v_{O})/(n_{O}v_{H})$; $n_{H}$ \& $v_{H}$ and $n_{O}$ \& $v_{O}$ are the densities \& velocities of, respectively, hydrogen and oxygen. Note, that for thermal motions $v_{O}/v_{H}$ is equal to $0.25$. If the ratio $\eta=\nu/N_{ph}$ stays constant for different grains, it is evident from Eq.(\ref{12}) that $t_{p}$ at the first approximation is $\approx 0.25\frac{\nu_{cr}^{2}}{\eta\gamma_{1}}t_{h}\frac{n_{H}}{n_{O}}$, i.e. it is independent of the number of active sites for $\nu>\nu_{cr}$. This is a new result, which was not foreseen in Paper II. This result means, for instance, that the estimates of $t_{p}$ obtained there for $\nu=\nu_{cr}$ are applicable to a wide range of grains with $\nu>\nu_{cr}$. Poisoning of active sites is essential for the Purcell alignment. A parameter that enters this theory is the ratio of $t_{x}$ given by Eq.(\ref{5}) to the time scale of paramagnetic relaxation $t_{mag}$ (see eq.~(58) in Purcell 1979, also see Roberge et al 1993). To provide sufficient alignment this ratio should be greater than unity (see fig.2 in Purcell 1979). To obtain this for standard values of the ISM parameters \footnote{It is argued in Paper II that these standard values may be misleading for particular regions of diffuse clouds, but we avoid discussing this issue here.} one needs to assume long-lived spin-up, i.e. $t_{L}\gg t_{d}$ (Spitzer \& McGlynn 1979) and therefore $t_{x}\sim t_{L}$. It is easy to check that $t_{mag}$ scales as $l^{2}$, where $l$ is a grain size and therefore on its own paramagnetic alignment favors small grains (see Johnson 1982). For $\nu>\nu_{cr}$ in the first approximation we have obtained that $t_{p}$ scales as $l^{0}$ (see Eq.~(12)). Note, that for $\nu<\nu_{cr}$ our arguments above are not applicable. The corresponding study in Paper II showed, that in this case not more than one active site is expected to be empty and the time of poisoning is proportional to $\nu t_{O} N_{ph}/m_{h}$ for $m_{h}\ll N_{ph}$. The number of hops for $\nu<\nu_{cr}$ is the ratio of time scale of the arrival of a hydrogen atom $t_{H}$ and the time scale of the hop of the oxygen atom $t_{h}$. Therefore both $t_{O}$ and $m_{h}$ scale as $l^{-2}$, while both $N_{ph}$ and $\nu$ scale as $l^{2}$. As a result the time of poisoning scales as $l^{4}$, which makes long-lived spin-up more probable for large grains. In general, the life time of Purcell's rockets $t_{L}$ is the minimal of the following time scales: $t_{p}$ and the time scale of accreting one monolayer of refractory material. In diffuse clouds, as a rule, accreting of ice mantles is suppressed (see Tanaka et al 1990). It is also believed, that hydrogenated nitrogen and carbon within diffuse clouds do not form mantles either. As the abundance of heavier elements is negligible, it is natural to assume that in diffuse clouds $t_{L}$ is controlled by poisoning of active sites. Above we disregarded photodesorption. The characteristic time for this process $t_{pd}$ scales as $l^{0}$ for smooth grains. The situation $t_{pd}\gg t_{p}$ corresponds to that studied in Paper II. If $t_{pd}< t_{p}$, photodesorption removes molecules blocking the access to active sites quicker that these sites are being poisoned. Thus a long-lived spin-up is called into being. As a result the following qualitative picture emerges. For small grains poisoning dominates, i.e. $t_{pd}> t_{p}$, and therefore the spin-up is short-lived. The time of poisoning for $\nu<\nu_{cr}$ grows with the size as $l^{4}$ and for some size becomes equal to $t_{ph}$. Starting from this size we deal with long-lived spin-up; $t_{L}$ grows with the size until $\nu=\nu_{cr}$. After the critical number of active sites is exceeded, $t_{L}$ is stabilized at the attained level. In other words, marginal alignment is expected for small grains and good alignment for large grains; this corresponds to observations (see Kim \& Martin 1994, Kim \& Martin 1995). Using both results obtained above and in Paper II we may attempt to write equations for the dinamics of the number of active sites. Indeed, while photodesprption, cosmic ray bombardment etc. create (or recover) active sites, poisoning removes them. The rate of active site creation per unit area $A$ may be assumed constant, while the rate of poisoning is different for $\nu<\nu_{cr}$ and $\nu>\nu_{cr}$. If $\nu<\nu_{cr}$ the results obtained in Paper II indicate, that the rate of poisoning can be approximated by $CN_{ph}^{-1}$, where $C= 10^{12}\exp(-E_{h}/kT_{s})$ s$^{-1}$, while this rate is of the order of $CN_{ph}^{-1}\nu^{2}\nu_{cr}^{-2}$ if $\nu>\nu_{cr}$. Therefore for $\nu<\nu_{cr}$ the number of the active sites changes as \begin{equation} \frac{{\rm d}\nu}{{\rm d} t}=A-C\frac{1}{N_{ph}}, \label{nu1} \end{equation} while for $\nu>\nu_{cr}$ the following equation is valid: \begin{equation} \frac{{\rm d}\nu}{{\rm d} t}=A-C\frac{1}{N_{ph}}\frac{\nu^{2}}{\nu_{cr}^{2}}. \label{nu2} \end{equation} Evidentely, Eq.~(\ref{nu1}) envisages a linear change of the number of active sites. If the rate of desorption is greater than poisoning, the number of active sites will increase linearly with time, unless all the possible active sites are invoked over grain surface, provided, that this number is less than $\nu_{cr}$) or alternatively the number of active sites is reached $\nu_{cr}$. In the opposite case when the rate of desorption is less than the rate of poisoning the number of active sites decreases with time untill all the active sites disappear. If $\nu>\nu_{cr}$ the solution of the Eq.~(\ref{nu2}) is as follows: \begin{equation} \nu=\nu_{cr}\sqrt{\frac{AN_{ph}}{C}}\frac{1+B\exp(-\frac{2}{\nu_{cr}} \sqrt{\frac{AC}{N_{ph}}}t)}{1-B\exp(-\frac{2}{\nu_{cr}}, \sqrt{\frac{AC}{N_{ph}}}t)}. \label{nu3} \end{equation} where B is related to the number of active sites $\nu_{in}$ at $t=0$ in the following way: \begin{equation} B=\frac{\nu_{in}-\sqrt{\frac{A}{C}}}{\nu_{in}+\sqrt{\frac{A}{C}}}. \end{equation} Eq.~(\ref{nu3}) testifies, that the number of active sites tend to stabilize at the level $\nu_{cr}\sqrt{AN_{ph}/C}$. Therefore for $A$ a bit larger than $C/N_{ph}$, the number of active sites is expected to be of the order $\nu_{cr}$ in correspondence with a qualitative conclusion reached in Paper II. Unfortunately, in our ignorance of many parameters involved, it is extremely difficult to quantify this otherwise luring picture. First of all, we have rather uncertain knowledge of the density of active sites as well as of other parameters of grain surface. Then, photodesorption presents another problem, as the rates obtained in some laboratory experiments (see Bourdon, Prince \& Duley 1982) are very low. We are not sure either whether the photodesorption is driven by UV quanta only or also by 3 $\mu$m radiation as it is claimed in Williams et al (1992). In view of this ambiguities, we need to treat the picture above as a conjecture only. We believe, that further research in the field will test this conjecture. An interesting feature of Eq~(\ref{12}) that it shows that $t_{p}$ is proportional to the number of sites of physical adsorption. Therefore fractal grains with large surface area should correspond to larger $N_{ph}$. The alignment of fractal grains was studied in Lazarian (1995d), where the excess of the physical area for such grains over that for smooth grains was invoked in order to improve the efficiency of alignment. This study indicated a possibility of a substantial improvement of alignment, provided that molecules blocking the access to active sites had sufficient mobility. This assumption seems to be a shortcoming of the latter study. However, it is possible to show that the dependences for alignment measure on the fractal dimension obtained in Lazarian (1995d) are valid without this assumption just as the result of the dependence of $t_{p}$ on $N_{ph}$. Another worry in Lazarian (1995d) was, that for fractal grains the majority of H$_{2}$ formation events may take place within narrow pores, and this may suppress suprathermal rotation. Our arguments above show, that the sites within grain pores are likely to be poisoned and therefore H$_{2}$ formation should take place mainly over open surfaces of grains, which are kept clean e.g. due to photodesorption. At the same time, surfaces within pores provide oxygen with more space for its random walk hopping. To summarize, we have shown, that the Purcell (1979) mechanism analogously to the Mathis (1986) one favors large grains. To tell these two mechanisms it is advantageous to study dependences of degree of alignment on grain temperatures. Our discussion above indicates, that $t_{p}$ is proportional to $t_{h}$, which varies exponentially with temperature. Therefore the Purcell alignment, unlike the Mathis one should be very sensitive to changes of grain temperature. Above we assumed that initially the concentration of active sites over grain surface is high. Although this corresponds to the modern picture of grain chemistry (see Tielens \& Allamandola 1987, Buch \& Zhang 1991), an alternative approach to the problem is also possible. Indeed, if the surface density of active sites may be less than $10^{-5}$ cm$^{2}$ and than only grains larger than $10^{-5}$ cm are likely to have at least one active site (Lazarian 1994b).\footnote{I suggested this possibility, but did not treat it seriously untill it got support from John Mathis.} This idea requires further study, but here we would like to point out, that the case of high and low density of active sites can be distinguished by temperature dependence. It is easy to see, that if the density of active sites is low, while nascent H$_{2}$0 molecules are ejected on formation, the temperature dependence of the Purcell alignment is suppressed. \section{Other possibilities of alignment} {\bf Alignment of helical grains.} Grains studied above were symmetric. However, real grains may have helicity. In fact, grains in Purcell (1975, 1979), that rotate suprathermally due to variations of the accommodation coefficient or photoelectric emission coefficient are helical. What was omitted in the above study is that the two aforementioned effects can produce alignment even without paramagnetic relaxation. For instance, radiation is, as a rule, anisotropic and this may influence the rotation caused by photoelectric emission. There is a substantial difference when a helical grain is subjected to a isotropic bombardment of atoms or photons and when it is subjected to a flux. In the former case, there is no difference between H$_{2}$ formation and other causes of suprathermal rotation; with high degree of accuracy it is possible to assume, that the regular torque acts only along the axis of major inertia (see Spitzer \& McGlynn 1979). If, however, a helical grain is subjected to a radiative or corpuscular flux, the component of torque perpendicular to the axis of major inertia can cause precession, which may significantly alter the alignment. An interesting example of helical grains was discussed in Dolginov \& Mytrophanov (1976). There, grains subjected to regular torques due to the difference in scattering of left- and right- circular polarized photons were considered. It was found, that if grains, either due to their chemical composition or due to their shape, have different cross-sections for left- and right-hand polarized quanta (see Dolginov \& Silantev 1976) their scattering of unpolarized light results in the spin-up. The efficiency of this process is maximal for wavelengths $\sim l$. For twisted prolate grains, increments of grain angular momentum can be shown \cite{dm} to be of the order of $0.25\Phi_{p}\hbar l^{2}(n_{r}-1)^{4}$, where $\Phi_{p}$ is the flux of photons and $n_{r}$ is the refraction coefficient ($\sim 1.3$). Then the characteristic angular velocity is \begin{equation} \Omega\approx\frac{1}{8}\frac{\Phi_{p}(n_{r}-1)^{4}\hbar}{nm_{a}v_{a}l^{2}}, \end{equation} which is of the order $10^{7}$~s$^{-1}$ for the typical concentration of photons in the ISM $\sim 3\cdot 10^{9}$~cm$^{-2}$~s$^{-1}$ and may become much higher in the vicinity of bright sources. For optimal shape suggested in Dolginov \& Mytrophanov (1976) the scattering efficiency is sufficient to account for the alignment over vast regions, but one may expect the grains to have relatively small deviations towards the optimal form. Such grains are believed to be a natural product of evolution (Mathis 1990). However, numerical studies of the interaction of irregular grains with radiation by Bruce Draine (private communication) showed, that grains that do not resemble helicies can exibit high efficiency of spin-up due to scattering of radiation. These studies can shed the light on the problem, whether grain spin-up arising from light scattering is an exeption or a rule. Note, that such a spin-up is expected to be long-lived one. Indeed, it is not the surface, but grain volume, that should be altered substantially to cause a crossover event. Such a change is expected to take place over timescale much longer as compared with the time of gaseous damping, which means that the spin-up should be long-lived. For such a spin-up the alignment is nearly perfect just due to paramagnetic relaxation. However, helical grains can be aligned mechanically on the timescale much shorter than the one relevant to the paramagnetic alignment. Mechanical alignment of helical grains is a complex problem and we are going to subject this issue to scrutiny in our next paper. Here we just refer to the study in Dolginov \& Mytrophanov (1976) where it was shown, that grains asymptotically tend to align perfectly with their helicity axes along the direction of magnetic field, if the time of precession is much less than that of alignment, and along the flux direction, if the opposite is true. In their paper Dolginov \& Mytrophanov considered a tilted oblate grain which helicity axis coincided with the major inertia axis and a tilted prolate grain where the two axis were perpendicular. The conclusion reached in Dolginov \& Mytrophanov (1976) was that the two species subjected to a flux should be aligned orthogonally. However, the above study omits the influence of internal relaxation. Due to this effect a prolate grain initially rotating about its axis of minimal inertia will in a short period of time turn to rotate about the axis of major inertia. In other words, long axis of prolate and oblate grains should be aligned in the same direction. To obtain quantitative results applicable to the ISM and circumstellar regions one needs to estimate the relative importance of this particular type of suprathermal rotation. Estimates in Purcell (1979) show that, as a rule, angular velocity of rotation due to H$_{2}$ formation should dominate for the ISM. Therefore, it seems unlikely that alignment due to grain helicity is more important in diffuse clouds in comparison with the one discussed in Sect.~3. However, the alignment of helical grains may be essential in the vicinity of bright sources.\\ {\bf Alignment due to radiation fluxes.} Above we discussed both mechanical and paramagnetic alignment of suprathermally rotating grains. One may wonder whether alignment can be caused by absorption of quanta. Indeed, due to absorption of an individual quantum a grain gains $\hbar$ angular momentum. This physical process is invoked in Harwit (1970). However, it was shown later in Purcell \& Spitzer (1971) that if a grain is being subjected to a radiation flux, the drift produced by the radiation pressure entails Gold alignment, which is far more efficient than the Harwit one. It is not difficult to show, that the ratio of squared increments of angular momentum for the Harwit process $\langle( \triangle J_{H}^{2})\rangle$ to the one for the Gold process $\langle( \triangle J_{G}^{2})\rangle$ is negligible even for rather special conditions of alignment discussed in Aitken et al (1985). Note, that in the aforementioned study the Gold alignment was neglected and therefore a conclusion, that Harwit mechanism is efficient was reached. We claim, that this inference can be true only, if the Gold alignment is suppressed. The latter happens, for instance, if charged grains are trapped by close loops of magnetic field and cannot be accelerated under radiation pressure. In any case, such an inefficient process as a Harwit one, is unlikely to influence the dynamics of suprathermally rotating grains. This, however, does not mean, that radiation cannot influence alignment of suprathermally rotating grains in a way other than through differential scattering of circular polarized quanta. For instance, we may suggest a mechanism based on photodesorption. Indeed, if suprathermal rotation is due to H$_{2}$ formation, we may refer to the toy model discussed in Sect.~3, but use a radiation flux instead of gaseous one. If this radiation flux desorbs molecules blocking the access to active sites, the life time of Purcell's rockets and therefore $t_{L}$ will be different for $\bot$ and $\parallel$ orientation of our grain. It is easy to see, that this alignment is most efficient if $t_{L}\gg t_{d}$. However, in some cases, e.g. for hot and small grains discussed in Purcell \& Spitzer (1971), the radiation dumping may dominate the gaseous one. In this case the grain temperature should depend on the grain orientation and $t_{L}\gg t_{d}$ is not required for efficient alignment. However, a detailed study of these processes is beyond the scope of our present paper.\\ {\bf Alignment due to grain magnetic moments}. To make our discussion more complete, we need to mention the alignment due to grain magnetic moments. One of the causes of these moments can be charge (Martin 1971, Draine \& Salpeter 1979, Draine \& Sutin 1987) over rotating grains (Rowland effect). The torques that act on a rotating charged grain were studied in Davis \& Greenstein (1951), along with their famous prediction of paramagnetic alignment. However, the influence of the ambient gas was omitted in their analysis, and thus the conclusion that the `distribution must be completely independent of the field' was made. This was quoted in other sources and therefore may still cause confusion. Consider an isolated rotating charged grain in magnetic field. Its rotation creates a magnetic moment ${\bf \cal M}= \Sigma_{i}{\rm e}(2c)^{-1}a_{i}^{2}{\bf J}({\it I})^{-1}$, where $a_{i}$ is the distance from an elementary charge over the grain surface to the axis of rotation. This causes grain precession in the magnetic field according to the Larmour equation \begin{equation} \frac{{\rm d}{\bf J}}{{\rm d}{\it t}}= {\bf \cal M}\times{\bf B}= \omega_{g} {\bf J}\times\frac{{\bf B}}{| {\bf B}|} \label{56} \end{equation} where $\omega_{g}=\sum \frac{{\rm e}a_{i}^{2}B}{3cI}$, which is quite close to the grain Larmour frequency $\omega_{L}$. In the course of this precession the angle between ${\bf \cal M}$ and ${\bf B}$ does not change, i.e. the precession by itself cannot change the energy of the system. The calculations in Davis \& Greenstein (1951) reflect this fact. However, the energy changes due to the interaction of grains with surrounding gas. Indeed, the impact in the direction of the Larmour precession arising from grain interaction with an atom results in the torque, that cause the precession around the axis which is perpendicular to the magnetic field. Eq.~(\ref{56}) shows that the angle between ${\bf \cal M}$ and ${\bf B}$ increases. If the impact is directed against the direction the Larmour precession, this angle decreases.\footnote{Note, that in reality we assume a uniform distributions of atomic impacts and subdivide the impacts as having component either in the direction of the precession or opposite to it.} The collisions of the first type disalign ${\bf \cal M}$ and ${\bf B}$, while the collisions of the second type align the two vectors. The alignment effect prevails as the impacts against precession are stronger than those in the direction of precession. In our arguments above we used the fact, that the Larmour precession influences the direction, but not the magnitude of the vector $\bf J$. Therefore the magnitude of ${\bf \cal M}$ (i.e. $|{\bf \cal M}|$) is not altered by the Larmour precession. In fact, in terms of grain interaction with magnetic field, the grain behaves as a magnetic dipole and the alignment due to the Rowland effect is similar to the alignment of magnetic dipoles. Such an alignment is not complete due to random torques acting upon grains.\footnote{Thermal fluctuations within grain material can randomize alignment as well, but the influence of such fluctuations for suprathermally rotating grains is negligible.} These random torques determine the effective temperature $T$ and this temperature influences the equilibrium distribution of ${\bf \cal M}$. If the suprathermal rotation is caused by the variations of the accommodation coefficient (see Purcell 1979), this temperature will be the mean of the grain and gas temperatures, provided that gas-grain collisions are inelastic. In the case of the suprathermal rotation caused by H$_{2}$ formation, $T$ will be of the order of $E_{H2}/k$, provided that every hydrogen atom that hits grain surface leaves it as a part of H$_{2}$ molecule. Indeed, if the spin-up is very long, then grains should behave as magnetic moments whose motion is disturbed by stochastic torques due to H$_{2}$ formation.\footnote{Note, that we are speaking about averaged $J^{2}$ components of the torque.} These magnetic moments are peculiar in a sense, that the randomization of their precession requires time much greater than usual damping time. However, if the life time of Purcell's rockets exceeds this time, the treatment in mu228rv2 should be applicable. In the extreme case of the short spin-up, when all torques are essentially stochastic one may introduce the averaged magnetic moment and study precession of this moment in the magnetic field. Our arguments should be applicable and the damping of the precession should occur over the timescale of the order of rotational damping time. The equilibrium distribution of grain axes of major inertia in magnetic field $B$ is proportional to $\exp(- \mbox{$\cal E$}/(kT))$, where ${\cal E}=-{\bf \cal M}{\bf B}$. It is easy to see that \begin{equation} \frac{\mbox{$\cal E$}}{kT}\approx \frac{\rm e}{2c}l^{2}\frac{\omega_{T}B}{kT} \sim\frac{\omega_{L}}{\omega_{T}} \label{54} \end{equation} for a thermally rotating grain and $\mbox{$\cal E$}(kT)^{-1}\approx \omega_{L}\omega_{r}\omega_{T}^{-2}$ for a suprathermally rotating grain, $\omega_{T}$ is the frequency of rotation corresponding to the temperature $T$, $\omega_{r}$ is the frequency of the suprathermal rotation and $\omega_{L}$ is the Larmour frequency. In fact, grain alignment due to the Rowland effect is in no way different from the alignment of ferromagnetic grains with high permanent magnetization discussed in Spitzer \& Tukey (1951). The difference is of quantitative nature; ferromagnetic grains have much larger magnetic moments and therefore much more susceptible for such an alignment. However, even with ferromagnetic grains the alignment was shown to be not efficient for diffuse clouds \footnote{ It is possible to show that Spitzer \& Tukey alignment may not be negligible in dark clouds, where both grain and gas temperatures are low, while magnetic fields $> 10^{-4}$ are commonplace. However, such grains are not expected to rotate suprathermally and thus we do not discuss them there.} (Spitzer \& Tukey 1951). Therefore alignment due to the Rowland effect is negligible for the majority of the foreseeable cases. This a fortiori true, as paramagnetic grains obtain much larger magnetic moments through the Barnett, rather than through Rowland effect. Assuming that $\omega_{T}\approx 10^{5}$~s$^{-1}$, $B \approx 3\times 10^{-6}$~G, $\omega_{r}\approx 10^{4}\omega_{T}$, we still fall short by approximately six orders of magnitude to produce any measurable alignment due to this effect. To summarize, magnetic moments irrespectively of their nature (the Rowland effect, the Barnett effect, permanent magnetization) do produce alignment, but this alignment is negligible for suprathermally rotating grains in diffuse clouds. \section{Conclusions} Several mechanisms of alignment of suprathermally rotating grains were discussed and it was shown that I. Paramagnetic alignment stays the strongest candidate to account for grain alignment over vast regions of the diffuse clouds. Its modification suggested in Purcell (1979) depends on subtle processes over grain surfaces and this provides an opportunity to test it. II. Alignment of grains due to the cross section effect that was introduced in this paper can also be important for the diffuse ISM. This mechanism, similar to the paramagnetic one, tends to align long grain axis perpendicular to magnetic field lines, provided that the grain drift is caused by Alfv\'{e}nic perturbations. III. Radiation can drive alignment in the vicinity of bright sources even in the absence of supersonic drift. Such an alignment is more likely to arise from differential photodesorption, although in some particular cases anisotropic radiation field can also provide the alignment of ''helical`` grains. Therefore additional care is required in interpreting the corresponding polarimetric data. IV. Magnetic moments of grains both arising from the Barnett effect and due to their charge produce alignment, but the expected degree of alignment is negligible. {\bf Acknowledgement}\\ The initial impetus for this work came from comments by B.~Draine. I also would like to thank him for a subsequent discussion of the results and for numerous suggestions, remarks and comments, which considerably improved the paper. This work also owes much to my discussions with A.~Goodman, R.~Kulsrud, P.~Myers, M.~Rees, L.~Spitzer and D.~Williams. The research was partially supported by Isaak Newton Scholarship from IoA, University of Cambridge (UK), Visiting Fellowship at CfA, and NASA grant NAG5 2773.
\section{Introduction} \eqnum{0} \hspace*{\parindent} The unification of all fundamental interactions including gravity is one of the final goals of theoretical physics. The electromagnetic and weak interactions were unified by Weinberg and Salam, and grand unified theories (GUTs) have been proposed as a unification model of three fundamental interactions. In this unification scheme, all interactions are described by gauge fields. Furthermore, supersymmetry is proposed to unify interaction (bosons) and matter (fermions), and gravity could be included in a supergravity theory. Such unified theories are sometimes discussed in higher-dimensions. Then, the idea of superstring arises as an approach to the unification of all interactions and particles, giving ``theory of everything". Then we have recognized that gravity is one of the most important keys for unification. In order to understand the role of gravity in a fundamental unified theory, it is necessary and helpful to study concrete physical phenomena with strong gravity such as cosmology or black holes. We find new aspects of gravity and other fundamental fields through such theoretical studies, which might give us hints about unification. In the effective theories derived from the higher-dimensional unified theories\cite{Callan}, the dilaton field couples to other known matter fields. The coupling constant depends on the fundamental unified theory and the dimensionality of spacetime. Thus it is important to study how the coupling affects physical phenomena. The coupling plays some important roles in black hole physics\cite{GM} as well as in cosmology\cite{Dcos}. In this paper, we further study effects of a dilaton field on black hole physics, and in particular we will analyze the role of Hawking's quantum radiation. We consider the model with a dilaton field coupled to a U(1) gauge field, i.e., the Einstein--Maxwell--dilaton theory. The action is \begin{equation} S = \frac{1}{16 \pi} \int \mbox{d}^{4} x \sqrt{- g} \left[ R - 2 \left( \nabla \phi \right)^{2} - \mbox{e}^{- 2 \alpha \phi} F^{2} \right] \; , \label{eqn:dilaction} \end{equation} where $\phi$ and $F_{\mu\nu}$ are a dilaton field and U(1) gauge field, respectively, with coupling constant $\alpha$\cite{unit}. For a superstring, we may also include an axion field $H_{\mu\nu\rho}$. The action is then \begin{equation} S = \frac{1}{16 \pi} \int \mbox{d}^{4} x \sqrt{- g} \left[ R - 2 \left( \nabla \phi \right)^{2} - \mbox{e}^{- 2 \phi} F^{2} - \frac{1}{12} \mbox{e}^{- 4 \phi} H^2 \right] \; . \label{eqn:senaction} \end{equation} The action (\ref{eqn:dilaction}) reduces to the Einstein--Maxwell theory when the coupling constant $\alpha = 0$. The black hole solution for this case is the well known Kerr--Newman family. The case of $\alpha = \sqrt{3}$ corresponds to the 4-dimensional effective model reduced from the 5-dimensional Kaluza--Klein theory. The action (\ref{eqn:senaction}), in which the dilaton coupling constant $\alpha$ to the U(1) gauge field is unity, is a bosonic part of the low energy limit of superstring theory. The exact spherically-symmetric dilatonic black hole solution with arbitrary coupling constant $\alpha$ is known\cite{GM,GHS}. They have some interesting thermodynamical properties, which are not found in the conventional charged (Reissner-Nordstr\"{o}m) black hole. In particular, the temperature of the black hole in the extreme limit depends drastically on $\alpha$. If $\alpha < 1$, the temperature of the black hole vanishes in the extreme limit, as does that of the Reissner-Nordstr\"om black hole. On the other hand, the temperature of the extreme black hole with $\alpha > 1$ diverges. For $\alpha = 1$, it is a non-zero finite value. This new thermodynamical property implies that the emission rate of Hawking quantum radiation may be completely different, depending on the coupling constant. We expect that when $\alpha > 1$, the emission rate diverges in the extreme limit, because the temperature diverges. The black hole may evaporate very rapidly. However, it was pointed out \cite{HW} that for $\alpha >1$, the effective potential, over which created particles travel to an asymptotically flat region to evaporate, grows infinitely high in the extreme limit. Hence, Holzhey and Wilczek expected that the emission rate will be suppressed to a finite value. Since these two features are competing processes in Hawking radiation, it is not trivial to decide whether or not the emission rate from the extreme dilatonic black holes with $\alpha > 1$ diverges. Thus, we analyze the emission rates numerically under the assumption that the charge is conserved, and clarify what happens in the extreme limit. This is the main purpose of the present paper. In addition to the spherically symmetric black hole, rotating dilatonic black holes also have similar thermodynamical properties\cite{HH,Sen,KM}. We considered superradiance around the rotating dilatonic black holes in the previous paper\cite{KM} and showed that there is a critical value ($\alpha \sim 1$) beyond which the emission rate changes drastically. In this paper, we extend our analysis to include the role of the temperature, i.e., Hawking quantum radiation, which automatically includes a superradiant effect, and discuss the fate of rotating black holes due to the evaporation process. We only know two exact rotating black hole solutions for the action (\ref{eqn:dilaction}): Kerr--Newman ($\alpha = 0$) and Kaluza--Klein solution ($\alpha = \sqrt{3}$)\cite{FZB}. In the superstring case ($\alpha = 1$), Sen\cite{Sen} derived a rotating black hole solution for the action(\ref{eqn:senaction}). This solution is not exactly the same as those in the model (\ref{eqn:dilaction}), but we expect that the existence of the axion field will not drastically change the dependence of the emission rate on the dilaton coupling. Hence, we analyze these three black hole solutions and compare their emission rates. Besides these exact solutions, we consider an approximate solution of slowly rotating black holes with arbitrary coupling $\alpha$\cite{HH,Shiraishi1}. All rotating dilatonic black holes reduce to the Kerr solution when their charges vanish. We expect that the coupling constant dependence is most noticeable when the black hole is highly charged. We therefore analyze Hawking radiation from highly charged black holes. As we know, a charged black hole generally emits its charge at a high rate in the process of evaporation and so its charge will be quickly lost, unless the charge is conserved. Because we are now interested in the effect of the dilaton coupling on the emission rates, we first assume that the charge is conserved, which is true for a central charge. We then study the discharge processes to see how it is affected by the dilaton coupling. This paper is organized as follows. In the next section, we study Hawking radiation for a spherically symmetric dilatonic black hole and analyze the behaviour of the emission rate in the extreme limit. The emission rates from rotating black holes are presented in the section 3. It is assumed that the charge of the black hole is conserved. We discuss the evolution and fate of these black holes. The effects of the discharge process are considered by calculating superradiance in the spherically symmetric black hole in the section 4. Finally, we give our conclusions and remarks in the final section. \section{Hawking Radiation from Spherically Symmetric Dilatonic Black Holes} \eqnum{0} \hspace*{\parindent} We first consider spherically symmetric dilatonic black holes. In this case we know the exact solution with arbitrary coupling constant $\alpha$ \cite{GM,GHS}, which is given by \begin{eqnarray} \mbox{d}s^{2} & = & -{\Delta(\rho) \over {R^2(\rho)}} \: \mbox{d}t^{2} + \frac{{R^2(\rho)}}{\Delta(\rho)} \: \mbox{d}\rho^{2} + { {R}^{2}(\rho)} (\mbox{d} \theta^{2} + \sin^{2}\theta \: \mbox{d} \varphi^{2} ) \: , \nonumber \\ A_{t} & = & \frac{Q}{\rho}, \; \; \; \phi = \frac{\alpha}{1+\alpha^{2}} \ln \left(1 - \frac{\rho_{-}}{\rho} \right) \; , \label{eqn:nonrotsol} \end{eqnarray} where \begin{equation} \Delta (\rho) = \left( \rho - \rho_{+} \right) \left( \rho - \rho_{-} \right), \; \; \; R (\rho) = \rho \left( 1 - \frac{\rho_{-}}{\rho} \right)^{\alpha^{2}/(1 +\alpha^{2})} , \nonumber \\ \end{equation} and $A_t$ is the $t$-component of the gauge potential $A_\mu$. The outer and `inner' horizons $\rho_{+}$ and $\rho_{-}$ are given by the mass $M$, the electric charge of the black hole $Q$ and $\alpha$ as \begin{equation} \rho_{\pm} = \frac{(1+\alpha^2) (M \pm \sqrt{ M^{2} - \left( 1-\alpha^{2} \right) Q^{2}})}{\left(1 \pm \alpha^{2} \right)}. \label{eqn:parametersph} \end{equation} $\rho = \rho_{-}$ is the curvature singularity for $\alpha \neq 0$. The maximum value of the charge is $Q_{\mbox{\scriptsize max}} \equiv \sqrt{1+\alpha^2} M$. When $|Q| = Q_{\mbox{\scriptsize max}}$, $\rho_{+}$ and $\rho_{-}$ coincide, and we call it an extreme black hole. However, it has to be emphasized that when $\rho_{+} = \rho_{-}$, a naked singularity appears at $\rho = \rho_{+}$ and the area of black hole vanishes for $\alpha \neq 0$, and it is therefore not a black hole solution\cite{GM}. \\ \hspace*{\parindent} The temperature $T$ of the black hole is given as \begin{equation} T = \frac{1}{4 \pi \rho_{+}} \left(1-\frac{\rho_{-}}{\rho_{+}} \right)^{(1-\alpha^2)/(1+\alpha^2)} . \end{equation} It possesses an interesting property\cite{GM}. When $\alpha < 1$, $T$ in the extreme limit vanishes, whereas it diverges in the case of $\alpha > 1$, and has the non-zero finite value $1 / 8 \pi M$ (as the Schwarzschild black hole) for $\alpha = 1$. \\ \hspace*{\parindent} Here we consider a neutral and massless scalar field which does not couple to the dilaton field\cite{Shiraishi2}, which is described by the Klein--Gordon equation \begin{equation} \Phi_{,\mu}^{~;\mu} = 0 \; . \label{eqn:KGeq} \end{equation} The energy emission rate of Hawking radiation is given\cite{Hawking} by \begin{equation} \frac{\mbox{d} M}{\mbox{d} t} = - \frac{1}{2 \pi} \sum_{l,m} \int_{0}^{\infty} \frac{\omega \: (1 - |A|^2)}{\exp \left[ \omega / \: T \right] - 1} \: \mbox{d} \omega , \label{eqn:dMsph} \end{equation} where $l$, $m$ are the angular momentum and its azimuthal component, $\omega$ is the energy of the particle, and $|A|^2$ is a reflection coefficient in a scattering problem for the scalar field $\Phi$. The Klein-Gordon equation (\ref{eqn:KGeq}) in this black hole spacetime can be made separable, by setting \begin{equation} \Phi = \frac{\chi(\rho^*)}{R(\rho)} \: S(\theta) \: \mbox{e}^{\mbox{\scriptsize i} m \varphi} \: \mbox{e}^{- \mbox{\scriptsize i} \omega t} . \label{eqn:Phiputform} \end{equation} Then, Eq.(\ref{eqn:KGeq}) is reduced to the Legendre equation for $S(\theta)$ and the radial equation, \begin{equation} \left[ \frac{\mbox{d}^2}{{\mbox{d} {\rho}^*}^{2}} + \omega ^{2} - {V}^2 (\rho) \right] \chi ({\rho}^*) = 0 , \label{eqn:KGsph} \end{equation} where \begin{eqnarray} V^2 (\rho) & \equiv & \frac{\Delta (\rho)}{R^{2} (\rho)} \left[ \frac{l(l+1)}{R^{2} (\rho)} + \frac{1}{R (\rho)} \frac{\mbox{d}}{\mbox{d} \rho} \left( {\Delta(\rho) \over {R^2(\rho)}} \frac{\mbox{d} R (\rho)}{\mbox{d} \rho} \right) \right] \: , \label{eqn:effpot} \\ \mbox{d} {\rho}^* & \equiv & \frac{ {R^2(\rho)}}{\Delta (\rho)} \; \mbox{d} \rho . \label{eqn:tortdef} \end{eqnarray} The reflection coefficient $|A|^2$ can be calculated by solving the wave equation (\ref{eqn:KGsph}) numerically under the boundary condition \begin{eqnarray} \chi & \rightarrow & \mbox{e}^{- \mbox{\scriptsize i} \omega \rho^*} + A \: \mbox{e}^{\mbox{\scriptsize i} \omega \rho^*} ~~~~~~ \mbox{as} ~~~ \rho^* \rightarrow \infty \: , \nonumber \\ \chi & \rightarrow & B \: \mbox{e}^{- \mbox{\scriptsize i} \omega \rho^*} ~~~~~~~~~~~~~~~ \mbox{as} ~~~ \rho^* \rightarrow - \infty \; . \label{eqn:boundcond} \end{eqnarray} \hspace*{\parindent} The dependence of the temperature $T$ on $\alpha$ might be expected to imply that the behaviour of Hawking radiation, which is thermal and has an emission rate proportional to $T^4$, is drastically affected by the dilaton coupling, particularly for $\alpha > 1$, for which the temperature $T$ diverges in the extreme limit. However, as Holzhey and Wilczek\cite{HW} pointed out, since the effective potential $V$ (\ref{eqn:effpot}) for $\alpha > 1$ grows infinitely high at the horizon in the extreme limit, the transmission probability $1 - |A|^2$ for particles to escape to infinity is suppressed. These two tendencies have opposite effects on Hawking radiation, and it is not clear whether or not the emission rate is actually suppressed. Here, we solve the wave equation (\ref{eqn:KGsph}) numerically to get the spectrum, and integrate Eq.(\ref{eqn:dMsph}). In this and subsequent calculations, we consider only the dominant modes with $l \leq 1$ since the contribution from higher angular momentum modes is suppressed by the centrifugal barrier. We integrate Eq.(\ref {eqn:dMsph}) numerically to $\omega_{\mbox{\scriptsize max}}$ ($\omega_{\mbox{\scriptsize max}} = 25 T$ for the present non-rotating case), which is justified since the spectrum is suppressed at the high energy regime by the exponential decay in the Planck distribution. \\ \hspace*{\parindent} To see how the emission rate varies as the black hole reaches to the extreme limit, we plot the emission rate, normalized by mass of the black hole $M$, against $Q / Q_{\mbox{\scriptsize max}}$ for five values of the coupling constant: $\alpha = 0$, $0.5$, $1$, $1.5$, $2$. It is shown in Fig.1. Here, we assume the charge of the black hole is positive, without loss of generality. \begin{figure} \singlefig{10cm}{dMsph.ps} \begin{figcaption}{fig:dMsph}{15cm} The emission rate for the non-rotating dilatonic black holes. \\ The charge $Q$ is normalized by $Q_{\mbox{\scriptsize max}}$ and the emission rate $-d M/d t$ is normalized by $M$. Each line corresponds to (1):$\alpha = 0$, (2):$\alpha = 0.5$, (3):$\alpha = 1$, (4):$\alpha = 1.5$, and (5):$\alpha = 2$, respectively. \end{figcaption} \end{figure} In this figure, we see that although the emission rates for each value of $\alpha$ coincide at $Q =0$, since the black hole solution, with any $\alpha$, is identically the Schwarzschild spacetime for $Q = 0$, the difference becomes large as the charge increases. In particular, the emission rate for $\alpha > 1$ blows up near the extreme limit. This means that the divergence of the temperature $T$ in the extreme limit overcomes that of the potential $V$. Furthermore, the emission rate (\ref{eqn:dMsph}) of the extreme black hole with $\alpha < 1$ is exactly zero because the temperature vanishes, and that for $\alpha = 1$ is non-zero but finite, as we see in the figure. Therefore, we may conclude that the behaviour of the emission rate in the extreme limit changes drastically at the value of $\alpha = 1$, as we naively expect from the behaviour of the temperature, despite the effect of the potential barrier. We may also speculate that nearly extreme black holes with $\alpha > 1$ are not stable objects. \section{Hawking Radiation from Rotating Dilatonic Black Holes} \subsection{Rotating Dilatonic Black Holes} \eqnum{0} \hspace*{\parindent} Next, we consider Hawking radiation from rotating black holes. In the rotating case, we know only two exact solutions in the model (\ref{eqn:dilaction}): the Kerr--Newman ($\alpha = 0$) and the Kaluza--Klein ($\alpha = \sqrt{3}$) solution\cite{FZB}. Besides these two, in the $\alpha = 1$ case, an exact rotating black hole solution is derived by Sen\cite{Sen} in the model (\ref{eqn:senaction}). We first summarize these exact solutions and their thermodynamical properties. \\ \hspace*{\parindent} Firstly, the Kerr--Newman black hole solution is expressed as \begin{eqnarray} \mbox{d}s^2 & = & - \frac{\Delta - a^2 \sin^2\theta}{\Sigma} \: \mbox{d}t^2 - \frac{2 a \sin^2\theta \left( r^2 + a^2 - \Delta \right)}{\Sigma} \: \mbox{d}t \mbox{d}\varphi \nonumber \\ & & + \frac{\left( r^2 + a^2 \right)^2 - \Delta a^2 \sin^2\theta}{\Sigma} \sin^2\theta \: \mbox{d}\varphi^2 + \frac{\Sigma}{\Delta} \, \mbox{d}r^2 + \Sigma \: \mbox{d}\theta^2 \: , \nonumber \\ A_t & = & \frac{Q r}{\Sigma} , \; \; \; A_\varphi = - \frac{a Q r \sin^2\theta}{\Sigma} \; , \label{eqn:KNsol} \end{eqnarray} where the functions $\Delta$ and $\Sigma$ are defined by \begin{equation} \Delta \equiv r^2 - 2 M r + a^2 + Q^2 \; , \; \; \; \Sigma \equiv r^2 + a^2 \cos^2\theta \; . \end{equation} The coordinate $r$ and $\rho$ in the previous section are related by $r = \rho - \rho_{-}$ in the spherically symmetric case. The temperature $T$ and the angular velocity $\Omega_{H}$ are given by \begin{eqnarray} T ~ & = & \frac{1}{2 \pi} \frac{\sqrt{M^2 - a^2 - Q^2}}{r_{\scriptscriptstyle H}^2 + a^2} \: , \label{eqn:TKN} \\ \Omega_{H} & = & \frac{a}{r_{\scriptscriptstyle H}^2 + a^2} \label{eqn:OmHKN} \: , \end{eqnarray} where \begin{equation} r_{\scriptscriptstyle H} = M + \sqrt{M^2 - a^2 - Q^2} \end{equation} is the horizon radius, $M$, $Q$, and $J = M a$ are the mass, the charge, and the angular momentum of the black hole, respectively. \\ \hspace*{\parindent} Secondly, the Kaluza--Klein black hole solution is derived by a dimensional reduction of the boosted 5-dimensional Kerr solution to four dimensions\cite{FZB,GW}. It is given by \begin{eqnarray} \mbox{d}s^{2} & = &- \frac{\Delta -a^2 \sin^2 \theta}{B \Sigma} \mbox{d}t^{2} - 2 a \sin^{2} \theta \frac{1}{ \sqrt{1-v^{2}}} \frac{Z}{B} \mbox{d}t \mbox{d} \varphi \nonumber \\ & & + \left[ B \left( r^{2} + a^{2} \right) + a^{2} \sin^{2} \theta \; \frac{Z}{B} \right] \sin^{2} \theta \; \mbox{d} \varphi^{2} + \frac{B \Sigma}{\Delta} \mbox{d}r^{2} + B \Sigma \, \mbox{d} \theta^{2} \: , \nonumber \\ A_{t} & = &\frac{v}{2 \left( 1-v^{2} \right) } \frac{Z}{B^{2}}, \; \; \; A_{ \varphi } = \; - a \sin^{2} \theta \frac{v}{2 \sqrt{1-v^{2}}} \frac{Z}{B^{2}}, \; \; \; \phi = \; - \frac{\sqrt{3}}{2} \ln B \; , \label{eqn:rotKKBHsol} \end{eqnarray} where \begin{equation} \Delta \equiv r^{2} - 2 \mu r + a^{2}, \; \; \; \Sigma \equiv r^{2} + a^{2} \cos^{2} \theta, \; \; \; Z \equiv \frac{2 \mu r}{\Sigma}, \; \; \; B \equiv \left( 1 + \frac{v^{2} Z}{1-v^{2}} \right)^{\frac{1}{2}} . \label{eqn:funcdefKK} \end{equation} The physical mass $M$, the charge $Q$, and the angular momentum $J$ are expressed by the parameters $v$, $\mu$, and $a$, as \begin{equation} M = \mu\left[1 + {v^2 \over 2(1-v^2)} \right] \; , \; \; \; Q = {\mu v \over 1-v^2} \; , \; \; \; J = {\mu a \over \sqrt{1-v^2}} \; . \label{eqn:physKK} \end{equation} The horizon radius is given by \begin{equation} r_{\scriptscriptstyle H} = \mu + \sqrt{ \mu^{2} - a^{2}} \; , \label{eqn:horizonKK} \end{equation} and then the regular horizon exists if \begin{equation} \mu^2 \geq a^2 \; , \label{eqn:horizoncond} \end{equation} and this condition may be rewritten as \begin{equation} \left({J \over M^2}\right)^2 \leq \frac{1}{4} \left[ 2- 10 \left({Q \over M} \right)^2 - \left({Q \over M} \right)^4 + 2 \left(1+ 2 \left({Q \over M} \right)^2 \right)^{3/2} \right] \; . \label{eqn:parameterrangeKK} \end{equation} The parameter range of the condition (\ref{eqn:parameterrangeKK}) is shown in Fig.2. It should be noted again that the solutions with $|Q| = Q_{\mbox{\scriptsize max}} \: (= 2 M)$ are not black hole solutions and these points are indicated by small circles in Fig.2. \begin{figure} \singlefig{10cm}{ParameterRange.ps} \begin{figcaption}{fig:ParameterRange}{15cm} The parameter ranges of three types of black hole. \\ The extreme lines are shown in $Q/M$--$J/M^2$ plane for (1):the Kerr--Newman, (2):the Sen, and (3):the Kaluza--Klein black holes. The region inside each line guarantees a regular event horizon, except the points denoted by a small circle where a naked singularity appears. \end{figcaption} \end{figure} \\ \hspace*{\parindent} As for the thermodynamical properties of this black hole, we find that the temperature $T$ and the angular velocity $\Omega_{H}$ are given as \begin{eqnarray} T ~ & = & \frac{\sqrt{1 - v^2}}{2 \pi} \frac{\sqrt{\mu^2 - a^2}}{r_{\scriptscriptstyle H}^2 + a^2} \label{eqn:TKK} \; , \\ \Omega_{H} & = & \frac{a \sqrt{1 - v^{2}}}{r_{\scriptscriptstyle H}^{2} + a^{2}} \; . \label{eqn:OmKK} \end{eqnarray} The temperature $T$ in the limit of $|Q| \rightarrow Q_{\mbox{\scriptsize max}}$ for the non-rotating black hole diverges, as was pointed out in the previous section. However, the temperature $T$ of the extreme rotating black hole ($\mu = |a|$) vanishes from Eq.(\ref{eqn:TKK}). When we take the limit $|Q| \rightarrow Q_{\mbox{\scriptsize max}}$, keeping the black hole extreme with $J \neq 0$ (whereas $J \rightarrow 0$ in the limit), the limiting value is still zero, and different from that of the non-rotating case. That is, the temperature is discontinuous at $|Q| = Q_{\mbox{\scriptsize max}}$ where a naked singularity appears. A similar feature is found in the behaviour of $\Omega_{H}$. If we take a limit $|Q| \rightarrow Q_{\mbox{\scriptsize max}}$, $\Omega_{H}$ of a rotating black hole diverges, whereas $\Omega_{H}$ of a non-rotating black hole is zero. The fact that $J$ vanishes while $\Omega_{H}$ diverges in the limit $|Q| \rightarrow Q_{\mbox{\scriptsize max}}$ is understood by observing that the area of the black hole vanishes in that limit. Those features are illustrated in Fig.3. \begin{figure} \begin{center} \segmentfig{11cm}{TKK.ps}{(a)} \vskip 1cm \segmentfig{10cm}{OmKK.ps}{(b)} \end{center} \begin{figcaption}{fig:thermKK}{15cm} The thermodynamical behaviour of the Kaluza--Klein black hole.\\ The behaviour of (a):the temperature $T$, and (b):the angular velocity $\Omega_{H}$ is depicted on $Q/M$--$J/M^2$ plane ($Q \geq 0$, $J \geq 0$). The constant angular momentum lines are drawn in solid lines for $J = 0$, $0.05 M^2$, and $0.2 M^2$. \end{figcaption} \end{figure} \\ \hspace*{\parindent} Thirdly, the Sen black hole\cite{Sen}, which is a solution in the action (\ref{eqn:senaction}), is expressed as \begin{eqnarray} \mbox{d}s^2 & = & -\frac{\Delta - a^2 \sin^2\theta}{\Sigma} \: \mbox {d}t^2 - \frac{4 \mu r a \cosh^2\beta \sin^2\theta}{\Sigma} \: \mbox {d}t \mbox{d} \varphi \nonumber \\ & & + \frac{\Sigma}{\Delta} \: \mbox{d} r^2 + \Sigma \: \mbox{d} \theta^2 + \frac{\Lambda}{\Sigma} \sin^2\theta \; \mbox{d} \varphi^2 \: , \nonumber \\ A_{t} & = & \frac{1}{\sqrt{2}} \frac{\mu r \sinh2\beta}{\Sigma} \: , \; \; \; A_{\varphi} = - \frac{a}{\sqrt{2}} \sin^2\theta \; \frac{\mu r \sinh2\beta}{\Sigma} \: , \nonumber \\ \phi & = & - \frac{1}{2} \: \ln \frac{\Sigma}{r^2 + a^2 \cos^2\theta} \: , \; \; \; B_{t \varphi} = 2 a \sin^2\theta \; \frac{\mu r \sinh^2\beta} {\Sigma} \; , \label{eqn:Sensol} \end{eqnarray} where the functions $\Delta$, $\Sigma$, and $\Lambda$ are defined by \begin{eqnarray} \Delta & \equiv & r^2 - 2 \mu r + a^2 , \; \; \; \Sigma \equiv r^2 + a^2 \cos^2\theta + 2 \mu r \sinh^2\beta \: , \nonumber \\ \Lambda & \equiv & \left( r^2 + a^2 \right) \left( r^2 + a^2 \cos^2 \theta \right) + 2 \mu r a^2 \sin^2\theta \nonumber \\ & & + 4 \mu r \left( r^2 + a^2 \right) \sinh^2\beta + 4 \mu^2 r^2 \sinh^4\beta \; . \label{eqn:Senfuncdef} \end{eqnarray} \hspace*{\parindent} The antisymmetric two rank tensor $B_{\mu\nu}$ generates the axion field $H_{\mu\nu\rho}$, together with $A_{\mu}$, by \begin{equation} H_{\mu\nu\rho} = \left( \partial_{\mu} B_{\nu\rho} - 2 A_{\mu} F_ {\nu\rho} \right) + \mbox{[cyclic permutations]} \; . \end{equation} The mass $M$, the charge $Q$, and the angular momentum $J$ are given by parameters $\mu$, $\beta$, and $a$ as \begin{equation} M = \frac{\mu}{2} \left( 1+\cosh2\beta \right), \; \; \; Q = \frac{\mu} {\sqrt{2}} \sinh2\beta, \; \; \; J = \frac{a \mu}{2} \left( 1+\cosh2\beta \right) \: , \end{equation} and the horizon radius is given by the same equation as Eq.(\ref {eqn:horizonKK}). The condition for the solution to be a black hole is also the same as Eq.(\ref{eqn:horizoncond}), which is now rewritten as \begin{equation} |J| \leq M^2 - \frac{Q^2}{2} \: . \label{eqn:parameterrangeSen} \end{equation} The parameter range of the condition (\ref{eqn:parameterrangeSen}) is also shown in Fig.2. \\ \hspace*{\parindent} This black hole has similar thermodynamical properties to the Kaluza--Klein solution. The temperature $T$ and the angular velocity $\Omega_{H}$ of this black hole are given by \begin{eqnarray} T & = & \frac{\sqrt{\left( 2 M^2 - Q^2 \right)^2 - 4 J^2}}{4 \pi M \left[ 2 M^2 - Q^2 + \sqrt{\left( 2 M^2 - Q^2 \right)^2 - 4 J^2} \right]} \; , \nonumber \\ \Omega_{H} & = & \frac{J}{M \left[ 2 M^2 - Q^2 + \sqrt{\left( 2 M^2 - Q^2 \right)^2 - 4 J^2} \right]} \; , \end{eqnarray} and these quantities are discontinuous at $|Q| = Q_{\mbox{\scriptsize max}} \: (= \sqrt{2} M)$, although they never diverge but approach finite values. The behaviour of these quantities is shown in Fig.4. \begin{figure} \begin{center} \segmentfig{11cm}{TSen.ps}{(a)} \vskip 1cm \segmentfig{10cm}{OmSen.ps}{(b)} \end{center} \begin{figcaption}{fig:thermSen}{15cm} The thermodynamical behaviour of the Sen black hole.\\ The same figures as Fig.3 are depicted for the Sen black hole. \end{figcaption} \end{figure} \\ \hspace*{\parindent} These discontinuities indicate that the emission rate of Hawking radiation may be completely different from that of the non-rotating case. In addition to the thermal effect of the temperature and the effective potential, which we considered in the previous section, new effects by the angular velocity are important in the rotating cases: in other words, superradiance. \subsection{Hawking Radiation of Rotating Black Holes} \hspace*{\parindent} Here, we discuss the radiation from rotating dilatonic black holes when the black hole charge is conserved. Hereafter, we can assume that $Q$ and $J$ are positive without loss of generality. The Klein--Gordon equation (\ref{eqn:KGeq}) for the neutral massless scalar field is separated into the spheroidal equation \begin{equation} \left[ \frac{1}{\sin\theta} \frac{\mbox{d}}{\mbox{d} \theta} \left( \sin\theta \frac{\mbox{d}}{\mbox{d} \theta} \right) - \left\{ a^2 \omega^2 \sin^2\theta + \frac{m^2}{\sin^2\theta} \right\} \right] S (\theta) = - \lambda S(\theta) \label{eqn:Spheroidaldil} \end{equation} and the radial equation \begin{equation} \left[ \frac{\mbox{d}^{2}}{\mbox{d} r^{* 2}} + \left(\omega - m \Omega (r) \right)^2 - V^2(r) \right] \chi \left( r^* \right) = 0 , \label{eqn:radialeqn} \end{equation} by setting \begin{equation} \Phi = \frac{\chi(r^*)}{R(r)} \: S(\theta) \: \mbox{e}^{\mbox {\scriptsize i} m \varphi} \mbox{e}^{- \mbox{\scriptsize i} \omega t} . \end{equation} Here the tortoise coordinate $r^*$ is defined by \begin{equation} \mbox{d} r^* \equiv \frac{R^2(r)}{\Delta(r)} \mbox{d} r \; . \end{equation} The functions $\Omega$, $R$, and $V$ are defined for the Kerr--Newman black hole as \begin{displaymath} \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \Omega(r) \equiv (2 M r -Q^2) \frac{a}{R^4(r)} \: , \; \; \; \; R^2(r) \equiv \left. \Sigma \right|_{\theta=0} = r^2 + a^2 \: , \end{displaymath} \begin{equation} V^2(r) \equiv {\Delta (r) \over R^2 (r) } \left\{ \frac{\lambda }{R^{2}(r)} + \frac{1}{R(r)} \frac{\mbox{d}}{\mbox{d} r} \left[ \frac{\Delta (r)}{R^2(r)} \frac{\mbox{d} R(r)}{\mbox{d}r} \right] - {m^2 a^2 \over R^6 (r) } (r^2 + a^2 - Q^2 + 2 M r) \right\} , \label{eqn:KGKNfncdef} \end{equation} for the Kaluza--Klein black hole as \begin{displaymath} \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \Omega(r) \equiv {2 \mu r \over \sqrt{1-v^2}}{a \over R^4(r)} \: , \; \; \; R^{2}(r) \equiv B \Sigma |_{\theta =0} = (r^{2} + a^{2}) \left( 1 + \frac{v^{2}}{1-v^{2}} \frac{2 \mu r}{r^{2} + a^{2}} \right)^{\frac{1}{2}} \: , \end{displaymath} \begin{equation} V^2(r) \equiv {\Delta (r) \over R^2 (r) } \left\{ \frac{\lambda }{R^{2}(r)} + \frac{1}{R(r)} \frac{\mbox{d}}{\mbox{d} r} \left[ \frac{\Delta (r)}{R^2(r)} \frac{\mbox{d} R(r)}{\mbox{d}r} \right] - {m^2 a^2 \over R^6 (r) } \left( r^2 +a^2 +{2 \mu r \over 1-v^2} \right) \right\} , \label{eqn:KGKKfncdef} \end{equation} and for the Sen black hole as \begin{displaymath} \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \! \Omega(r) \equiv 2 \mu r \cosh^2 \beta \frac{a}{R^4(r)} \: , \; \; \; R^2(r) \equiv \Sigma |_{\theta = 0} = r^2 + a^2 + 2 \mu r \sinh^2 \beta \; , \end{displaymath} \begin{equation} V^2(r) \equiv {\Delta (r) \over R^2 (r) } \left\{ \frac{\lambda}{R^2(r)} + \frac{1}{R(r)} \frac{\mbox{d}}{\mbox {d} r} \left[ \frac{\Delta(r)}{R^2(r)} \frac{\mbox{d} R(r)}{\mbox{d} r} \right] - \frac{m^2 a^2}{R^6(r)} \left( r^2 + a^2 + 2 \mu r \cosh 2\beta \right) \right\} . \label{eqn:KGSenfncdef} \end{equation} \hspace*{\parindent} The emission rates of energy and angular momentum\cite{Hawking} are given by \begin{eqnarray} \frac{\mbox{d} M}{\mbox{d} t} & = & - \frac{1}{2 \pi} \sum_{l, m} \int^{\infty}_{0} \frac{\omega \left( 1 - | A |^{2} \right)}{\exp \left[ \left( \omega - m \Omega_{H} \right) / \: T \right] - 1} \: \mbox{d} \omega \; , \nonumber \\ \frac{\mbox{d} J}{\mbox{d} t} & = & - \frac{1}{2 \pi} \sum_{l, m} \int^ {\infty}_{0} \frac{m \left( 1 - | A |^{2} \right)}{\exp \left[ \left( \omega - m \Omega_{H} \right) / \: T \right] - 1} \: \mbox{d} \omega \; . \label{eqn:radrot} \end{eqnarray} The reflection coefficient $|A|^2$ is calculated by solving the wave equation (\ref{eqn:radialeqn}) under the boundary condition \begin{eqnarray} \chi & \rightarrow & \mbox{e}^{- \mbox{\scriptsize i} \omega r^*} + A \: \mbox{e}^{\mbox{\scriptsize i} \omega r^*} ~~~~~~ \mbox{as} ~~~ r^* \rightarrow \infty \; , \nonumber \\ \chi & \rightarrow & B \: \mbox{e}^{- \mbox{\scriptsize i} \widetilde {\omega} r^*} ~~~~~~~~~~~~~~~ \mbox{as} ~~~ r^* \rightarrow - \infty \; , \end{eqnarray} where $\widetilde{\omega} \equiv \omega - m \Omega_{H}$. We integrate Eq.(\ref{eqn:radrot}) by setting the upper bound of the integration as $\omega_{\mbox{\scriptsize max}} = \max(25T, 1.5\Omega_{H} )$ for a rotating black hole. The eigenvalue $\lambda$ of the spheroidal equation (\ref {eqn:Spheroidaldil}) is calculated perturbatively\cite{Bouwkamp} as \addtolength{\abovedisplayskip}{0.2cm} \addtolength{\belowdisplayskip}{0.2cm} \addtolength{\jot}{0.2cm} \begin{eqnarray} \addtolength{\abovedisplayskip}{-0.2cm} \addtolength{\belowdisplayskip}{-0.2cm} \addtolength{\jot}{-0.2cm} \lambda & = & l(l+1) + \frac{1}{2} \left[ \frac{(2m-1) (2m+1)}{(2l-1) (2l+3)}+1 \right] a^2 \omega^2 + \frac{1}{2} \left[ \frac{(l-m-1) (l-m) (l+m-1) (l+m)}{(2l-3) (2l-1)^3 (2l+1)} \right. \nonumber \\ & & \left. -\frac{(l-m+1) (l-m+2) (l+m+1) (l+m+2)}{(2l+1) (2l+3)^3 (2l+ 5)} \right] a^4 \omega^4 + {\cal O} (a^6 \omega^6) . \end{eqnarray} This approximation is valid, since $a \omega < 1$ and the coefficient of each term is small for all the cases we analyzed, although $\omega M > 1$ in some instances. \\ \hspace*{\parindent} As we mentioned before, the coupling constant dependence of the temperature or the angular velocity is remarkable in highly charged black holes. Hence we analyze such cases. If the black hole has a large charge, it carries only a small angular momentum, as is seen from Fig.2. Because of this, we consider the black holes with a small angular momentum, which is fixed at $J = 0.01 M^2$, and vary the charge to see how the emission rates for each solution change in the extreme limit. The result is shown in Fig.5. The charge is normalized by the maximal value $Q_{\mbox{\scriptsize ex}}$ for the black hole with $J = 0.01 M^2$, which is a little less than $Q_{\mbox{\scriptsize max}}$ (See Fig.2). The values of $Q_{\mbox{\scriptsize ex}}$ are $0.999 \: Q_{\mbox{\scriptsize max}}$, $0.995 \: Q_{\mbox{\scriptsize max}}$, and $0.972 \: Q_{\mbox{\scriptsize max}}$ for the Kerr--Newman, the Sen and the Kaluza--Klein black holes, respectively. \begin{figure} \begin{center} \segmentfig{10cm}{dMrotex.ps}{(a)} \vskip 1cm \segmentfig{10cm}{dJrotex.ps}{(b)} \end{center} \begin{figcaption}{fig:radrot}{15cm} The emission rate of (a):the energy $-dM/dt$ and (b):the angular momentum $-dJ/dt$ for three types of rotating black holes. \\ $Q$ is normalized by $Q_{\mbox{\scriptsize ex}}$. Each line corresponds to (1):the Kerr--Newman, (2):the Sen, and (3):the Kaluza--Klein black holes, respectively. $Q_{\mbox{\scriptsize ex}}$ is $0.999 \: Q_{\mbox{\scriptsize max}}$, $0.995 \: Q_{\mbox{\scriptsize max}}$, and $0.972 \: Q_{\mbox{\scriptsize max}}$ for Kerr--Newman, for the Sen, and for the Kaluza--Klein black hole, respectively. \end{figcaption} \end{figure} In Fig.5, we see the Kaluza--Klein black hole radiates much more energy and angular momentum near the extreme limit than the Kerr--Newman and the Sen black holes. The behaviour of the energy emission rates (Fig.5 (a)) is very similar to that of non-rotating black holes, except in the vicinity of the extreme limit. This is because we have chosen a very small value for the angular momentum. However, in the Kaluza--Klein black hole, the emission rate drops a little near the extreme limit and does not diverge, so we find a different result from the non-rotating case. This is because the temperature of the rotating Kaluza--Klein black hole vanishes in the extreme limit whereas that of the non-rotating case is divergent in the same limit. \\ \hspace*{\parindent} There appears to be a critical value of the dilaton coupling constant at $\alpha \sim 1$, although we cannot give a definite critical value from our analysis of exact black hole solutions. However there is another way to investigate such a critical value in the extreme limit. The temperatures of rotating black holes vanish in the extreme limit, and it is known that Hawking radiation becomes purely superradiant\cite {Unruh}, that is, the emission rates (\ref{eqn:radrot}) are \begin{eqnarray} \frac{\mbox{d} M}{\mbox{d} t} & = & - \frac{1}{2 \pi} \sum_{l, m} \int^{m \Omega_{H}}_{0} \omega \left( | A |^{2} -1 \right) \mbox{d} \omega \; , \nonumber \\ \frac{\mbox{d} J}{\mbox{d} t} & = &- \frac{1}{2 \pi} \sum_{l, m} \int^ {m \Omega_{H}}_{0} m \left( | A |^{2} -1 \right) \mbox{d} \omega . \label{eqn:radrotsup} \end{eqnarray} \hspace*{\parindent} In the previous paper\cite{KM}, we analyzed superradiance of the rotating dilatonic black holes, which we will briefly summarize. To see how superradiance depends on the dilaton coupling constant, we considered the slowly rotating approximate solution with arbitrary coupling constant\cite{HH,Shiraishi1}, which is given by adding an angular momentum perturbation to the spherically symmetric solution (\ref{eqn:nonrotsol}), as well as the three exact solutions. This solution is expressed, in the same coordinates as the spherically symmetric solution in the section 2, as \begin{eqnarray} \mbox{d}s^{2} & = & -{\Delta(\rho) \over {R^2(\rho)}} \mbox{d}t^{2} + \frac{{R^2(\rho)}}{\Delta(\rho)} \mbox{d}\rho^{2} + { {R}^{2}(\rho)} (\mbox{d} \theta^{2} + \sin^{2} \theta \mbox{d} \varphi^ {2} ) - 2 a f(\rho) \sin^{2}\theta \: \mbox{d} t \mbox{d} \varphi \nonumber \\ A_{t} & = & \frac{Q}{\rho}, \; \; \; A_{\varphi} = - a \sin^{2} \theta \frac{Q}{\rho}, \; \; \; \phi = \frac{\alpha}{1+\alpha^{2}} \ln \left(1 - \frac{\rho_{-}}{\rho} \right) \; , \label{eqn:slowsol} \end{eqnarray} where \begin{eqnarray} \Delta (\rho) & \equiv & \left( \rho - \rho_{+} \right) \left( \rho - \rho_{-} \right), \; \; \; R (\rho) \equiv \rho \left( 1 - \frac{\rho_{-}}{\rho} \right)^{\alpha^ {2}/(1 +\alpha^{2})} \; , \nonumber \\ f (\rho) & \equiv & \frac{\left( 1+\alpha^{2} \right)^{2}}{\left( 1- \alpha^{2} \right) \left( 1-3\alpha^{2} \right) } \left({\rho \over \rho_{-}} \right)^2 \left( 1 - \frac{\rho_{-}}{\rho} \right)^{2\alpha^{2}/(1 +\alpha^{2})} - \left( 1 - \frac{\rho_{-}}{\rho} \right)^{(1-\alpha^{2})/(1+\alpha^{2})} \nonumber \\ & \times & \left( 1 + \frac{\left( 1+\alpha^{2} \right)^{2}}{\left( 1-\alpha^{2} \right) \left( 1-3\alpha^{2} \right) } \left({\rho \over \rho_{-}}\right)^2 + \frac{1+\alpha^{2}}{1-\alpha^{2}} \left({\rho \over \rho_{-}}\right) - \frac{\rho_{+}}{\rho} \right) \; , \end{eqnarray} and \begin{equation} \rho_{\pm} = \frac{(1+\alpha^2) (M \pm \sqrt{ M^{2} - \left( 1-\alpha^{2} \right) Q^{2}})}{\left(1 \pm \alpha^{2} \right)}, \; \; \; a = \frac{2 (1 + \alpha^2) J}{ (1 + \alpha^2) \rho_{+} + (1-\alpha^{2}/3) \rho_{-}} . \label{eqn:parameteraprx} \end{equation} This solution is valid only when the parameter $a$ is sufficiently small. Although $ f(\rho) $ seems to diverge at $ \alpha = 1/\sqrt{3} $, $ \alpha = 1 $ or $ \rho_{-} = 0 $, $f(\rho)$ approaches a finite limiting value when we expand this function around each point. \\ \hspace*{\parindent} The Klein-Gordon equation is now separated into the Legendre equation and the radial equation \begin{equation} \left[ \frac{\mbox{d}^2}{{\mbox{d} {\rho}^*}^{2}} + \left( \omega - m \Omega (\rho) \right)^{2} - {V}^2 (\rho) \right] \chi ({\rho}^*) = 0 \; , \label{eqn:radialslow} \end{equation} where \begin{eqnarray} \Omega (\rho) & \equiv & \frac{a f (\rho)}{ R^{2} (\rho)},\\ V^2 (\rho) & \equiv & \frac{\Delta (\rho)}{R^{2} (\rho)} \left[ \frac{l(l+1)}{R^{2} (\rho)} + \frac{1}{R (\rho)} \frac{\mbox{d}}{\mbox{d} \rho} \left( {\Delta(\rho) \over {R^2(\rho)}} \frac{\mbox{d} R (\rho)}{\mbox{d} \rho} \right) \right], \end{eqnarray} and ${\rho}^*$ is defined by \begin{equation} \mbox{d} {\rho}^* \equiv \frac{ {R^2(\rho)}}{\Delta (\rho)} \: \mbox{d} \rho \; . \label{eqn:tortoiseaprx} \end{equation} The emission rates by superradiance for this approximate black hole solution are shown in Fig.6, with the calculations using the three exact rotating black hole solutions. \begin{figure} \begin{center} \segmentfig{10cm}{dMrotsup.ps}{(a)} \vskip 1cm \segmentfig{10cm}{dJrotsup.ps}{(b)} \end{center} \begin{figcaption}{fig:radsup}{15cm} Superradiance from slowly rotating black holes. \\ (a) and (b) show the energy emission rate $-dM/dt$, and the angular momentum emission rate $-dJ/dt$, respectively. Each line corresponds to (1):$\alpha = 0$, (2):$\alpha = 0.5$, (3):$\alpha = 0.9$, (4):$\alpha = 1.1$, (5):$\alpha = 1.5$, (6):$\alpha = \sqrt{3}$, and (7):$ \alpha = 2$. In addition, we plot the results for three exact solutions (the circles for the Kerr--Newman, the squares for the Sen, and the triangles for the Kaluza--Klein black holes). The charge is normalized by $Q_{\mbox{\scriptsize max}}$. \end{figcaption} \end{figure} We find that the emission rate from the large coupling constant black holes blows up as the black hole approaches the extreme one, whereas with small $\alpha$, the emission rate remains quite small. These two types of behaviour are divided by a value of the coupling constant of about unity. \\ \hspace*{\parindent} Again we cannot determine the exact value of the critical coupling constant, for the following reason. As we mentioned before, this approximate black hole solution is valid only when the angular momentum is sufficiently small. In addition to this condition, there is another requirement that must be satisfied, namely, that the black hole charge should not be so large. This is found by observing that the maximally charged black hole ($Q = Q_{\mbox{\scriptsize max}}$) in the approximate solution can carry an angular momentum, while the exact solution cannot (e.g., consider the Kerr--Newman black hole). Quantitatively, the angular velocity $\Omega_{H} \equiv \Omega(\rho_{+})$ of the black hole is divergent in the extreme limit for $\alpha \geq 1/\sqrt{3}$ and vanishes for smaller coupling constants, but this critical value is derived from the approximate solution and may differ from the value of the exact solution, which we do not know. In fact, the angular velocity in the extreme limit of the Sen black hole, which is a solution of the model (\ref{eqn:senaction}) ($\alpha = 1$), is finite and non-zero, although this solution is obtained from a different action from the Kerr--Newman and the Kaluza--Klein black holes. The qualitative behaviour of the angular velocity in the approximate solution seems to follow that of the exact solution closely, although that of the temperature does not. Hence we may give a qualitative discussion of superradiance by using this approximate solution. \\ \hspace*{\parindent} We may conclude from the above that the critical coupling constant at which the behaviour of the superradiant emission changes exists and is about unity. As we have already shown, the behaviour of the emission rate by thermal radiation from the non-rotating black hole also changes at $\alpha = 1$. Naively speaking, Hawking radiation for the rotating black hole consists of two components, that is, thermal radiation and superradiance. So we naturally expect that the emission of Hawking radiation from rotating black holes is drastically changed at $\alpha \sim 1$. \subsection{The Fate of Dilatonic Black Holes} \hspace*{\parindent} The dependence of the emission on the coupling constant leads to a difference in the evolution of black holes by evaporation. To investigate the evolution of the three exact rotating solutions above, we describe the black hole state by a pair of quantities $(Q/M, J/M^2)$, and analyze their time variations, which are given by the emission rates as \begin{eqnarray} \frac{\mbox{d}}{\mbox{d} t} \left( \frac{Q}{M} \right) & = & - \frac{Q} {M^2} \frac{\mbox{d} M}{\mbox{d} t} \; , \label{eqn:evolQ} \\ \frac{\mbox{d}}{\mbox{d} t} \left( \frac{J}{M^2} \right) & = & \frac{1} {M^2} \frac{\mbox{d} J}{\mbox{d} t} - 2 \frac{J}{M^3} \frac{\mbox{d} M} {\mbox{d} t} \; , \label{eqn:evolJ} \end{eqnarray} for the three exact black hole solutions. We recognize these two quantities as a vector field on the $Q/M$--$J/M^2$ plane shown in Fig.2, and show it in Fig.7. \begin{figure} \begin{center} \segmentfig{5.7cm}{evolKN.ps}{(a)} \hspace{2cm} \segmentfig{7.66cm}{evolSen.ps}{(b)} \end{center} \vskip 2cm \begin{center} \segmentfig{16cm}{evolKK.ps}{(c)} \end{center} \begin{figcaption}{fig:EvolVec}{15cm} The evolution of three types of black hole. \\ Each figure represents (a):the Kerr--Newman, (b):the Sen, and (c):the Kaluza--Klein black hole. The arrow shows the direction and magnitude of the evolution of the black hole by Hawking evaporation at each point. The scale of the arrow is enlarged 2500 times. Although the arrows near $Q = Q_{\mbox{\scriptsize max}}$ are very small for the Kerr--Newman and the Sen black holes, those in the Kaluza--Klein black hole are considerably larger. \end{figcaption} \end{figure} Since we assume that the black hole charge is conserved, and the black holes lose mass energy, Eq.(\ref{eqn:evolQ}) is always positive, so $Q/ M$ increases and the black hole approaches the extreme state. In Fig.7, near the extreme lines, each vector points to a direction inside the extreme line, so the black hole does not evolve beyond the extreme line and eventually approaches the $Q = Q_{\mbox{\scriptsize max}}$ state. From the figure, we can see that the Kerr--Newman black hole stops its evolution as it approaches $Q = Q_{\mbox{\scriptsize max}}$ whereas the evolution of the Kaluza--Klein black hole is accelerated as $Q / M$ increases, and in particular, the evolution is very fast near $Q = Q_{\mbox{\scriptsize max}}$. This is because the emission rates of the Kaluza--Klein black hole near $Q = Q_{\mbox{\scriptsize max}}$ state are very large. As we mentioned before, the $Q = Q_{\mbox{\scriptsize max}}$ state is not a black hole solution and a naked singularity appears at this point. So it is indicated from our analysis that the Kaluza--Klein black hole evolves rapidly into a naked singularity. As we have already seen, the area of the Kaluza--Klein black hole vanishes as $Q \rightarrow Q_{\mbox{\scriptsize max}}$. This situation is quite similar to the evaporation of the Schwarzschild black hole, for which the area of the black hole vanishes and the emission rate increases infinitely large in the final stage, where a naked singularity might appear. The Sen black hole shows an intermediate behaviour between that of the Kerr-- Newman and the the Kaluza--Klein black holes. \section{Discharge of Dilatonic Black Holes by Superradiance} \eqnum{0} \hspace*{\parindent} So far, we have considered only the case where the charge of the black hole is conserved. Usually, however, black holes may create charged particles and lose their charge. In this section we study the discharge process by superradiance of a charged scalar field described by the equation of motion \begin{equation} \left[ \left( \nabla^\mu + \mbox{i} e A^\mu \right) \left( \nabla_\mu + \mbox{i} e A_\mu \right)- \mu^2 \right] \Phi = 0 \; , \label{eqn:KGq} \end{equation} where $e$ and $\mu$ are the charge and the rest mass of the particle, respectively. Shiraishi\cite{Shiraishi2} analyzed superradiance of a charged scalar field $\Phi$ coupled to the dilaton $\phi$ in the spherically symmetric dilatonic black hole. Here we do not consider such a coupling because we are only interested in the pure quantum properties of the dilatonic black hole, but not the extra effects on the quantum radiation, which come from a direct coupling between $\Phi$ and the dilaton field. \\ \hspace*{\parindent} The timescales of the loss of energy, angular momentum, and charge depend on the temperature $T$, the angular velocity $\Omega_{H}$, and the electric potential $\Phi_{H}$ in the Planck distribution of Hawking radiation as \begin{equation} \frac{1}{\exp\left[\left(\omega - m \Omega_{H} - e \Phi_{H}\right) / T \right]} \; . \end{equation} If the electric potential is large enough compared with the temperature and the angular velocity, the dominant component of the emission is that of the superradiant discharge process. In order to estimate how important the discharge process is in Hawking radiation, we calculate the superradiant emission rates in a spherically symmetric dilatonic black hole, in which the electric potential $\Phi_{H}$ is \begin{equation} \Phi_{H} = \frac{Q}{\rho_{+}} \; . \label{eqn:phiHdil} \end{equation} The horizon radius $\rho_{+}$ is given by Eq.(\ref{eqn:parametersph}). If superradiance is large compared to the emission calculated in the previous sections, where we assumed that the charge is conserved, the discharge process is important and should not be ignored, while, if it is small, the discharge process is not essential in Hawking radiation. \\ \hspace*{\parindent} The emission rates are \begin{eqnarray} \frac{\mbox{d} M}{\mbox{d} t} & = & - \frac{1}{2 \pi} \sum_{l , m , e} \int_{\mu}^{e \Phi_{H}} \omega \left( |A|^2 - 1 \right) \mbox{d}\omega \label{eqn:dMRNsuper} \; , \\ \frac{\mbox{d} Q}{\mbox{d} t} & = & - \frac{1}{2 \pi} \sum_{l , m , e} \int_{\mu}^{e \Phi_{H}} e \left( |A|^2 - 1 \right) \mbox{d}\omega \: , \label {eqn:dQRNsuper} \end{eqnarray} where the reflection coefficient $|A|^2$ is obtained by solving the radial wave equation \begin{equation} \left[ \frac{\mbox{d}^2}{{\mbox{d} {\rho}^*}^{2}} + \left( \omega - e \frac{Q}{\rho} \right)^{2} - \mu^2 \frac{\Delta(\rho)}{R^2(\rho)} - {V}^2 (\rho) \right] \chi ({\rho}^*) = 0 \; , \label{eqn:Radialdis} \end{equation} which is derived by setting in the same way as Eq.(\ref {eqn:Phiputform}), under the boundary condition of \begin{eqnarray} \chi & \rightarrow & \mbox{e}^{- \mbox{\scriptsize i} \omega \rho^*} + A \: \mbox{e}^{\mbox{\scriptsize i} \omega \rho^*} ~~~~~~ \mbox {as} ~~~ \rho^* \rightarrow \infty \; , \nonumber \\ \chi & \rightarrow & B \: \mbox{e}^{- \mbox{\scriptsize i} \widetilde {\omega} \rho^*} ~~~~~~~~~~~~~~~ \mbox{as} ~~~ \rho^* \rightarrow - \infty \; , \end{eqnarray} where $\widetilde{\omega}$ is now defined as $\widetilde{\omega} = \omega - e \Phi_{H}$, and the tortoise coordinate $\rho^*$, functions $R (\rho)$ and $\Delta(\rho)$, and the potential $V^2$ are the same as those in the section 2. Here we consider only the dominant mode of $l = 0$. \\ \hspace*{\parindent} The wave equation (\ref{eqn:Radialdis}) is not invariant under rescaling by the black hole mass $M$, in contrast to the case of the massless field considered in the previous sections. The first and the last terms in the bracket in Eq.(\ref{eqn:Radialdis}) are roughly proportional to $M^{-2}$, whereas the second and the third terms are independent of the mass scale. This results in that the transmission probability $|A|^2 - 1$ depends explicitly on the mass of the black hole. Hence we have to calculate the emission rates for each mass scale and analyze the mass dependence of the emission, in addition to the coupling constant dependence. \\ \hspace*{\parindent} First we consider the Planck mass black hole $(M = M_{\mbox {\scriptsize PL}})$. We show the emission rates in Fig.8 for four values of the coupling constant: $\alpha = 0$, $0.5$, $1$, $1.5$. $Q$ is now normalized by the mass of the black hole $M$, but not $Q_{\mbox{\scriptsize max}}$, because $Q$ itself is essential in this process, but not $Q/Q_{\mbox{\scriptsize max}}$. We set the particle mass $\mu = 0.001 M_{\mbox{\scriptsize PL}}$. \begin{figure} \begin{center} \segmentfig{10cm}{dMdisq1.ps}{(a)} \vskip 1cm \segmentfig{10cm}{dQdisq1.ps}{(b)} \end{center} \begin{figcaption}{fig:discharge1}{15cm} Discharge by superradiance from non-rotating black holes with mass $M = M_{\mbox{\scriptsize PL}.}$ \\ (a) and (b) show the energy emission rate $-dM/dt$, and the charge emission rate $-dQ/dt$ normalized by the Planck mass $M_{\mbox {\scriptsize PL}}$, respectively. Each line corresponds to (1):$\alpha = 0$, (2):$\alpha = 0.5$, (3):$\alpha = 1$, and (4):$\alpha = 1.5$. \end{figcaption} \end{figure} {}From this figure, we find that the emission rates are greater in the black hole with the smaller coupling constant, in contrast to the results of the previous two sections. In particular, emission from the highly charged black hole with larger coupling constant is very small. There are two reasons for this. One is the behaviour of the electric potential $\Phi_{H}$. From Eq.(\ref{eqn:phiHdil}), we can see the electric potential becomes smaller when the coupling constant $\alpha$ increases. The second reason is that the effective potential in Eq.(\ref{eqn:Radialdis}) is very high near the extreme limit for the black hole with $\alpha > 1$ and the transmission probability becomes much smaller, as in the previous cases. \\ \hspace*{\parindent} Now we analyze the dependence of the emission rates on the mass of the black hole. To see how the emission rate changes, we calculate the case of $M = 10 M_{\mbox{\scriptsize PL}}$ and show the result in Fig.9. \begin{figure} \begin{center} \segmentfig{10cm}{dMdisq10.ps}{(a)} \vskip 1cm \segmentfig{10cm}{dQdisq10.ps}{(b)} \end{center} \begin{figcaption}{fig:discharge10}{15cm} Discharge by superradiance from non-rotating black holes with mass $M = 10 M_{\mbox{\scriptsize PL}.}$ \\ (a) and (b) show the energy emission rate $-dM/dt$, and the charge emission rate $-dQ/dt$, respectively. Each line corresponds to (1):$\alpha = 0$, (2):$\alpha = 0.5$, (3):$\alpha = 1$, and (4):$\alpha = 1.5$. \end{figcaption} \end{figure} Comparison with Fig.8 shows that the emission rates generally increase when the mass increases. This tendency is clearer in the highly charged black holes with larger coupling constant. The dependence on the coupling constant is smaller than the case of $M = M_{\mbox{\scriptsize PL}}$. This is because the height of the effective potential, which is roughly proportional to $M^{-2}$, is effectively lower than that in the case of a Planck mass black hole. In particular, the emission of a highly charged black hole with large mass becomes insensitive to the coupling constant because the potential barrier gets small, compared with the case of a Planck mass-scale black hole where the potential is very high for the large coupling constant and the emission is suppressed near the extreme limit. Consequently, we expect that the coupling constant dependence of the emission rate will become smaller as we increase the mass of the black hole. \\ \hspace*{\parindent} For a black hole larger than $10 M_{\mbox{\scriptsize PL}}$, the numerical calculation becomes difficult because we have to deal with a very large scale black hole and a very small scale particle simultaneously. Fortunately, for a massive black hole with small charge, the $V^2$ term in Eq.(\ref{eqn:Radialdis}) is very small and can be neglected. Furthermore, the rest of the potential terms (the second and the third terms in the bracket) in Eq.(\ref{eqn:Radialdis}) vary very slowly, so we can use the W.K.B. approximation to calculate the transmission probabilities, as in Ref.\cite{Shiraishi2,Gibbons}. \\ \hspace*{\parindent} When the black hole mass $M$ is sufficiently large and $Q/M$ is small, the radial wave equation (\ref{eqn:Radialdis}) is approximated by \begin{equation} \left[ \frac{\mbox{d}^2}{{\mbox{d} {\rho}^*}^{2}} + \left( \omega - e \frac{Q}{\rho} \right)^{2} - \mu^2 \frac{\Delta(\rho)} {R^2(\rho)} \right] \chi ({\rho}^*) = 0 \; , \label{eqn:RadialWKB} \end{equation} and the transmission probability $|A|^2 - 1$ can be estimated from \begin{eqnarray} |A|^2 - 1 & = & \exp\left[ - 2 \int^{\rho_{2}^*}_{\rho_{1}^*} \sqrt{|W|} \: \mbox{d}\rho^ * \right] \nonumber \\ & = & \exp\left[ - 2 \int^{\rho_{2}}_{\rho_{1}} \sqrt{|W|} \: \frac{R^2} {\Delta} \: \mbox{d}\rho \right] \; , \label{eqn:WKBtrs} \end{eqnarray} where \begin{equation} W = \left( \omega - e \frac{Q}{\rho} \right)^{2} - \mu^2 \frac{\Delta (\rho)}{R^2(\rho)} \; , \end{equation} and $\rho_{1}^*$ and $\rho_{2}^*$ ($\rho_{1}^* < \rho_{2}^*$) are the corresponding tortoise coordinates to two roots $\rho_{1}$, $\rho_{2}$ of $W(\rho) = 0$. \\ \hspace*{\parindent} In the $\alpha = 0$ case, in which the black hole is described by the Reissner--Nordstr\"om solution and \begin{equation} W = \left( \omega - e \frac{Q}{\rho} \right)^2 - \mu^2 \left( 1 - \frac{2 M}{\rho} + \frac{Q^2}{\rho^2} \right) \; , \end{equation} Eq.(\ref{eqn:WKBtrs}) is integrated, giving \begin{equation} |A|^2 - 1 = \exp\left[ - 2 \pi \mu^2 \frac{e Q - \left( \omega - k \right) M}{k \left( \omega + k \right)} \right] \; , \label{eqn:transRN} \end{equation} where \begin{equation} k \equiv \sqrt{\omega^2 - \mu^2} \; . \end{equation} {}From this, we find Schwinger's formula for the emission rate $\mbox{d}Q/\mbox{d}t$ \begin{equation} \frac{\mbox{d} Q}{\mbox{d} t} \sim - \frac{e^4 Q^3}{\rho_{+}} \exp\left [-\frac{\pi \mu^2 \rho_{+}^2}{e Q} \right] \end{equation} in the small charge limit\cite{Gibbons}. \\ \hspace*{\parindent} We can also explicitly calculate the transmission probability in the superstring case ($\alpha = 1$), in which \begin{equation} W = \left( \omega - e \frac{Q}{\rho} \right)^2 - \mu^2 \left( 1 - \frac {\rho_{+}}{\rho} \right) \; . \end{equation} It gives exactly the same result as Eq.(\ref{eqn:transRN}). In addition, for the case of small charged black holes with arbitrary coupling constant, we can use the approximation \begin{equation} \frac{Q}{Q_{\mbox{\scriptsize max}}} \ll 1 \; , \end{equation} so $\rho_{+} \gg \rho_{-}$, and then \begin{equation} \frac{\Delta(\rho)}{R^2(\rho)} = \left( 1 - \frac{\rho_{+}}{\rho} \right) \left( 1 - \frac{\rho_{-}}{\rho} \right)^{(1 - \alpha^2)/(1 + \alpha^2)} \sim \; \left( 1 - \frac{\rho_{+}}{\rho} \right) \left( 1 - \frac{\tilde {\rho}_{-}}{\rho} \right) \; , \end{equation} where \begin{equation} \tilde{\rho}_{-} \: \equiv \: \frac{1 - \alpha^2}{1 + \alpha^2} \: \rho_{-} \; , \end{equation} and we find the same transmission probability as Eq.(\ref {eqn:transRN}). Hence, for the dilatonic black hole with a fixed mass and charge, the transmission probability of the particle with the same energy is hardly influenced by the coupling constant. As for the total emission rate, the black hole with the larger coupling constant emits a little bit less energy, because the energy range of the superradiant modes, i.e., $\mu \leq \omega \leq e \Phi_{H}$, becomes narrow as the coupling constant increases. When the charge of the black hole increases, the emission rate increases. In the extreme limit, the black hole with larger coupling constant can carry a larger charge. Hence we may expect that the nearly extreme black hole with a larger coupling constant emits larger energy than that with a smaller coupling constant. However, near the extreme limit for $\alpha > 1$, the W.K.B. approximation breaks down and the effective potential becomes very steep. As a result, emission may not increase so much. So we expect that the dependence of the emission on the coupling constant becomes smaller for a more massive black hole. This has been confirmed by our numerical calculations. \section{Conclusion and Discussion} \eqnum{0} \hspace*{\parindent} In summary, we first studied the evaporation of dilatonic black holes under the assumption that the black hole charge is conserved, and analyzed its dependence on the dilaton coupling constant. We found that the emission rate of the non-rotating black hole changes drastically at $\alpha = 1$, which is the value predicted by superstring theory. In the case of the coupling constant below unity, the emission rate vanishes in the extreme limit, while the black hole with $\alpha > 1$ emits a large amount of energy in the same limit, even though the potential barrier becomes infinitely high in this case. This means the effect of the temperature on the emission is stronger than that of the potential barrier. \\ \hspace*{\parindent} As for rotating black holes, the temperature is zero for the extreme black holes and the thermal emission also vanishes for all known exact black hole solutions. However, in the maximally charged limit $Q \rightarrow Q_{\mbox{\scriptsize max}}$ of the Kaluza--Klein black hole, while the angular momentum itself is still small, the angular velocity of the black hole becomes very large and the effect of superradiance becomes important. In superradiance, we also find the critical value of the coupling constant $\alpha \sim 1$, above which the emission rate increases rapidly as the black hole approaches the maximally charged state. Therefore, we may reasonably conclude that $\alpha \sim 1$ is the critical coupling constant together with the thermal component of the quantum radiation. \\ \hspace*{\parindent} As a result, a highly charged Kaluza--Klein black hole ($\alpha = \sqrt{3}$) is inevitably accelerated towards evaporation into a naked singularity. This situation is very similar to the final stage of the evaporation of the Schwarzschild black hole where the emission blows up and the area of the black hole vanishes. We expect that black holes with $\alpha > 1$ show a similar evaporation process to the Kaluza--Klein case, since the emission rates for such black holes are very large in the maximally charged limit. \\ \hspace*{\parindent} We have also considered the discharge process by calculating superradiance for non-rotating dilatonic black holes. If the mass of the black hole is on the Planck scale, the emission is suppressed for large coupling constants, compared with the Reissner--Nordstr\"om black hole ($\alpha = 0$), especially near the extreme limit. Hence, the effect of the discharge may not be so important for highly charged black holes with $\alpha > 1$. As the mass of the black hole increases, however, the dependence of the emission on the coupling constant becomes small and a black hole with any $\alpha$ will discharge efficiently. \\ \hspace*{\parindent} Holzhey and Wilczek\cite{HW} pointed out that, in the maximally charged limit of the dilatonic black holes, the thermodynamical interpretation breaks down. The solution of the maximally charged limit represents a naked singularity, and the higher order quantum effects will become important near this limit. This means that the black hole thermodynamics may deviate from the conventional approach, which is based on the semiclassical treatment of Hawking radiation. We should make some comments on this point. The problems related to this paper are: (1) The emission rate becomes very large, so we have to consider the backreaction of the quantum effects on the metric, (2) The area of the black hole vanishes in the maximally charged limit, which means we have to deal with a horizon radius smaller than the Planck scale, (3) To clarify the coupling constant dependence, we discuss the Planck mass-scale black hole. In order to study such problems properly, we may need quantum gravity. However, before investigating the full quantum theory, we first have to clarify the behaviour in the semiclassical regime. \vspace{0.5cm} {\bf Acknowledgment}\\ We would like to thank R. Easther for reading the paper carefully. This work was supported partially by the Grant-in-Aid for Scientific Research Fund of the Ministry of Education, Science and Culture (No. 06302021 and No. 06640412), and by Waseda University Grant for Special Research Projects. \vskip 2cm \baselineskip .2in
\section{Introduction} The existence of gravitational waves is an unambiguous prediction of the theory of general relativity \cite{Thorne87}. Yet, despite efforts originating in the early nineteen sixties, gravitational waves have not been detected directly. Nevertheless, observation of the binary pulsar PSR 1913+16 convinces us that gravitational waves do exist, and that they are correctly described by Einstein's theory \cite{Taylor}. That gravitational waves have not yet been detected on Earth is simply due to their incredible weakness: typical waves would produce in a bulk of matter a strain $\Delta L/L$, where $L$ is the extension of the matter, of order $10^{-21}$ \cite{Thorne87}. Needless to say, to measure this effect is a great challenge for experimentalists. \section{Interferometric detectors} There is reasonable hope that gravitational waves will be detected within the next ten years, thanks to a new generation of detectors which use interferometry to monitor the small displacements induced by the passage of a gravitational wave. Two groups are currently involved in building large-scale interferometers: the American LIGO team, and the French-Italian VIRGO team. The LIGO (Laser Interferometer Gravitational-wave Observatory) project \cite{LIGO} involves two detectors, to be built in Hanford, Washington, and in Livingston, Louisiana. Construction has begun at both sites. Each interferometer has an armlength of approximately 4 km. LIGO should be completed by the turn of the century. The VIRGO (so named after the galaxy cluster) project \cite{VIRGO} involves a single interferometer, to be built near Pisa, Italy, with an armlength of approximately 3 km. VIRGO should also be completed by the turn of the century. The basic idea behind interferometric detectors is the following \cite{Thorne87}: The interferometer is composed of two long (4 km for LIGO) vacuum pipes forming the letter {\bf L}. A laser beam is split in two at the corner of the {\bf L}, and is sent into each arm of the interferometer. Each beam then bounces off a mass which is suspended at each end of the {\bf L} (a mirror has been coated onto each mass). The light is finally recombined at the beam splitter, and its intensity is measured by a photodiode. When no gravitational wave is present at the interferometer, the length of each arm is so adjusted that when measured by the photodiode, the light's intensity is precisely zero (the recombined beams are arranged to be precisely out of phase). However, when a gravitational wave passes through the interferometer, the armlengths are no longer constant, and the recombined beams no longer precisely out of phase. More precisely, during the first half of its cycle the gravitational wave increases the length of one arm, and decreases the length of the other. During the second half cycle, the first arm is now shorter, and the second arm longer. The light's intensity therefore oscillates with the gravitational-wave frequency. The intensity is a measure of $\Delta L / L = h$, where $L$ denotes the interferometer armlength, and $h$ the gravitational-wave field. \section{Detector noise} Interferometers are subject to various sources of noise which limit the detector's sensitivity to gravitational waves. The relative importance of each source depends on the frequency at which the interferometer oscillates \cite{LIGO}. At low frequencies ($f < 10\ \mbox{Hz}$) the detector's sensitivity is limited by seismic noise, which is due to the Earth's seismic activity. At frequencies larger than 10 Hz the seismic noise can be eliminated with sophisticated isolation stacks; these fail at low frequencies. At high frequencies ($f > 100\ \mbox{Hz}$) the detector's sensitivity is limited by photon shot noise, which is due to statistical errors in the counting of photons by the photodiode. This source of noise can be reduced by increasing the laser power, or making use of ``light recycling'' \cite{LIGO}. At intermediate frequencies ($f$ between 10 Hz and 100 Hz) the noise is dominated by thermal noise, which is due to spurious motions of thermal origin. For example, the suspended masses are thermally excited and vibrate with their normal-mode frequencies; this evidently affects the recombined laser beam. Interferometers are therefore broad-band detectors, with good sensitivity in the range \cite{LIGO} \begin{equation} 10\ \mbox{Hz} < f < 1\ 000\ \mbox{Hz}. \label{1} \end{equation} The required sensitivity for full-scale interferometers is approximately $h_n \sim 10^{-22}$ at peak sensitivity --- a tall order. [The subscript $n$ stands for ``noise level''; we will define $h_n(f)$ precisely below. A plot of $h_n(f)$, appropriate for an interferometric detector with ``advanced'' sensitivity, is given in Fig.~1.] For comparison, we may mention that the Caltech 40 m prototype has already achieved $h_n \simeq 10^{-19}$ at peak sensitivity $(f=450\ \mbox{Hz})$. It is not implausible that improved technology and a factor of 100 in armlength will permit to reach the desired goal. \section{More about detector noise} \begin{figure}[t] \special{hscale=50 vscale=50 hoffset=60.0 voffset=-300.0 angle=0.0 psfile=fig1.ps} \vspace*{3.2in} \caption[Fig. 1]{Noise level in an interferometric detector with advanced sensitivity.} \end{figure} The detector noise can be measured when no gravitational wave is present at the interferometer, the typical situation. Then the detector output $s(t) = \Delta L(t)/L$ is given by noise alone: \begin{equation} s(t) = n(t), \label{2} \end{equation} where $n(t)$ represents the noise. The noise is a random process \cite{Reif}: the function $n(t)$ takes purely random values. Consequently, the noise can only be studied using statistical methods. In the presence of a gravitational wave, Eq.~(\ref{2}) must be replaced by $s(t) = h(t) + n(t)$, where $h(t)$ is the gravitational-wave field. The statistical properties of the noise can be determined by careful measurement. For example, the time average \begin{equation} \overline{n(t)} = \lim_{T\to\infty}\, \frac{1}{2T} \int_{-T}^{+T} n(t)\, dt \label{3} \end{equation} can be constructed. This mean value can then be subtracted from $n(t)$ and, without loss of generality, we can put $\overline{n} = 0$. Also from measurements, the noise's {\it autocorrelation function} $C_n(\tau)$ can be constructed: \begin{equation} C_n(\tau) = \overline{n(t) n(t+\tau)}; \label{4} \end{equation} $C_n(0)$ gives the mean squared deviation of the noise with respect to the mean value. In the following we will assume that the noise is {\it stationary}, in the sense that its statistical properties do not depend on time \cite{Reif}. This means, in particular, that the autocorrelation function does not depend explicitly on the origin of time $t$, but only on the variable $\tau$, as was expressed in Eq.~(\ref{4}). We shall also assume that the noise satisfies the {\it ergodic hypothesis}, so that time averages can be replaced with ensemble averages \cite{Reif}. Here, the noise is imagined to be drawn from a representative ensemble, and the probability that it takes a particular realization $n(t)$ is given by a specified probability distribution. The statistical properties of the noise then refer to this (infinite dimensional) distribution function. In full generality, the statistical properties of the noise can only be summarized by constructing all the higher moments $\overline{n \cdots n}$. If, however, the noise is assumed to be Gaussian, in the sense that its probability distribution function is an infinite dimensional Gaussian distribution \cite{Helstrom}, then the autocorrelation function contains all the information. Real detector noise is neither strictly stationary nor strictly Gaussian. However, on a timescale of hours, which is long compared with typical gravitational-wave bursts, the noise appears stationary to a good approximation \cite{LIGO}. And non-Gaussian components to the noise can be removed, to a large extent, by cross-correlating detector outputs from two widely separated interferometers \cite{LIGO}. It is therefore a satisfactory approximation to take the noise to be stationary and Gaussian. Under these assumptions the statistical properties of the detector noise are fully summarized by $C_n(\tau)$. It is convenient to work instead in the frequency domain, and to define \cite{Reif} the noise's {\it spectral density} $S_n(f)$ as \begin{equation} S_n(f) = 2 \int C_n(\tau) e^{2\pi i f \tau}\, d\tau. \label{5} \end{equation} The spectral density is defined for $f>0$ only; as $C_n(\tau)$ is a real and even function, the negative frequencies only duplicate the information contained in the positive frequencies. As $n(t)$ is dimensionless, the spectral density has dimensions of time. By multiplying $S_n(f)$ with the frequency and taking the square root (since the spectral density represents the mean squared noise), one obtains the {\it noise level} $h_n(f)$: \begin{equation} h_n(f) = \sqrt{ f S_n(f) }. \label{6} \end{equation} This gives the equivalent gravitational-wave amplitude which would make the interferometer oscillate at just the noise level \cite{Thorne87}; this quantity was introduced in the preceding section (see Fig.~1). \section{Coalescing compact binaries} Coalescing compact binaries, composed of neutron stars and/or black holes, are the most promising source of gravitational waves for interferometric detectors \cite{Thorne87,Schutz}. Consider the binary pulsar PSR 1913+16 \cite{Taylor}. This system consists of two neutron stars, each of $1.4\ M_\odot$, in orbital motion around each other. Its present orbital period $P$ is approximately 8 hours, corresponding to orbital separations of about $5\times 10^5$ times the total mass. (Here and throughout we use units such that $G=c=1$.) Its present eccentricity is approximately equal to 0.6. The binary's orbital period is observed to decay at a rate $dP/dt = -2 \times 10^{-12}$ corresponding precisely to a loss of energy and angular momentum to gravitational waves \cite{Taylor}. In a timescale of approximately $10^8$ years the orbital period will have decreased to less than a tenth of a second, corresponding to orbital separations smaller than one hundred times the total mass. In this time, the eccentricity will have been reduced to extremely small values (by the radiation reaction), so that the orbits are practically circular. The gravitational waves produced then have a frequency larger than 10 Hz, and the frequency keeps increasing as the system evolves. {\it During this late stage of orbital evolution, the gravitational waves sweep through the frequency bandwith of interferometric detectors, and thus become visible.} By the time the orbital separation becomes as small as a few times the total mass, the neutron stars begin to merge. The gravitational waves produced during the final merger cannot be detected by interferometric detectors, at least in the broad-band configuration described above \cite{foot1}: the gravitational-wave frequency is then larger than 1 000 Hz, for which the detector noise is large. Of course, it would be foolish to wait $10^8$ years in order for PSR 1913+16 to produce gravitational waves with appropriate frequency. Fortunately, interferometric detectors will be sensitive enough to monitor binary coalescences occurring in quite a large volume of the universe, approximately $10^7\ \mbox{Mpc}^3$ (corresponding to a radius of 200 Mpc \cite{LIGO}). It has been estimated \cite{Phinney} that as many as 100 coalescences could occur every year in such a volume (this includes coalescences of black-hole systems as well). This potentially large event rate is one of the factors that make gravitational waves from coalescing compact binaries especially promising. The other factor comes from the fact that compact binaries are extremely clean astrophysical systems. It can be estimated \cite{Kochanek} that tidal interactions between the two stars are completely negligible, up to the point where the objects are about to merge. In particular, no mass transfer occurs. The system can therefore be modeled, to extremely good accuracy, as that of two point masses with a limited number of internal properties (such as mass, spin, and quadrupole moment). The challenge in modeling coalescing compact binaries resides in formulating and solving the equations of motion and wave generation for a general relativistic two-body problem \cite{Damour}. \section{Waveform according to the quadrupole formula} At the crudest level, the gravitational-wave signal corresponding to an inspiraling binary system can be calculated by (i) assuming that the orbital motion is Newtonian (with the effects of radiation reaction incorporated) and (ii) using the standard quadrupole formula for wave generation \cite{MTW}. As motivated above, we may also assume that the orbits are circular. We define $h(t)$ to be the gravitational-wave signal. This is given by a linear combination, appropriate for interferometric detectors, of the two fundamental polarizations, $h_+$ and $h_\times$, of the gravitational-wave field. At this level of approximation, the signal is given by \cite{Thorne87} \begin{equation} h(t) = Q(\mbox{angles})\, ({\cal M}/r)\, (\pi {\cal M} f)^{2/3}\, \cos \Phi(t). \label{7} \end{equation} As expected, the signal decays as the inverse power of $r$, the distance to the source. In Eq.~(\ref{7}), $Q$ is a function of all the angles relevant to the problem: position of the source in the sky, orientation of the orbital plane, position and orientation of the detector on Earth. The parameter ${\cal M}$ is called the {\it chirp mass} and represents a particular combination of the masses, given by \begin{equation} {\cal M} = (m_1 m_2)^{3/5} / (m_1 + m_2)^{1/5}. \label{8} \end{equation} The waveform depends on the chirp mass only, and not on any other combination of the masses. The symbol $f$ represents the gravitational-wave frequency, which is equal to {\it twice} the orbital frequency. Because the system loses energy and angular momentum to gravitational waves, the frequency is not constant, but increases in time according to \cite{Thorne87} \begin{equation} \frac{df}{dt} = \frac{96}{5\pi {\cal M}^2} (\pi {\cal M} f)^{11/3}. \label{9} \end{equation} As a consequence, Eq.~(\ref{7}) shows that the amplitude of the signal, which is proportional to $(\pi {\cal M} f)^{2/3}$, also increases with time. A signal which increases both in frequency and in amplitude is known as a {\it chirp}, and this is the origin of the term ``chirp mass''. Finally, the phase function $\Phi(t)$ is given by \begin{equation} \Phi(t) = \int^t 2\pi f(t')\, dt'. \label{10} \end{equation} Because the frequency is not a constant, the phase accumulates nonlinearly with time. For a system of two neutron stars, the gravitational-wave signal undergoes approximately 16 000 oscillations as it sweeps through the frequency bandwith of an interferometric detector. The timescale for the frequency sweep is approximately 15 minutes. The orbital separation ranges from approximately 180 to 10 times the total mass $M = m_1 + m_2$, and the orbital velocity \begin{equation} v \equiv (\pi M f)^{1/3} \label{11} \end{equation} ranges from approximately 0.1 to 0.4. This indicates that relativistic corrections must be inserted in Eqs.~(\ref{7}) and (\ref{9}) in order to obtain a satisfactory degree of accuracy. \section{Matched filtering} How does one go about finding a gravitational-wave signal in a noisy data stream, when typically the signal is not very strong? And once the signal is found, how does one go about extracting the information it contains? For signals of precisely known form, one goes about this using the technique of {\it matched filtering} \cite{Wainstein}. Signals from inspiraling compact binaries, since they can be calculated with high precision, belong to this class. The basic idea behind matched filtering is to use our knowledge about the signal in order to go find it in the data stream, after the noisy frequencies have been filtered out. Suppose a signal of known form $h(t;\vec{\mu})$ is present in the data stream. Here, the vector $\vec{\mu}$ collectively denotes all the parameters characterizing the signal. In the case of inspiraling binaries, these would be the time of arrival, the initial phase, the distance to the source, the position angles, the chirp mass, and other parameters to be introduced in Sec.~12. The detector output is given by \begin{equation} s(t) = h(t;\vec{\mu}) + n(t), \label{12} \end{equation} where $n(t)$ is the noise, whose statistical properties are fully summarized by the spectral density $S_n(f)$, as discussed in Sec.~4. The first step in matched filtering \cite{CutlerFlanagan} is to pass $s(t)$ through a linear filter which removes the noisy frequencies. The idea here is to use our knowledge about the detector noise contained in $S_n(f)$ in order to discard that part of the detector output for which the detector noise is large. The output is imagined to be decomposed into Fourier modes according to $s(t) = \int \tilde{s}(f) e^{-2\pi i f t}\, df$; the filter suppresses the modes $\tilde{s}(f)$ for which $S_n(f)$ is large. In mathematical terminology, a linear filter is a linear operation on a function $s(t)$. This operation can always be written as \cite{Wainstein} \begin{equation} s(t) \to \int w(t-t') s(t')\ dt', \label{13} \end{equation} where $w(t-t')$ is the filter function. The filter which removes the noisy frequencies is the one such that \begin{equation} \tilde{w}(f) = \frac{2}{S_n(|f|)}; \label{14} \end{equation} the factor of 2 is conventional. The next step in matched filtering consists of computing the overlap integral between the filtered output (\ref{13}) and the known signal $h(t;\vec{\mu}_{\rm trial})$. The true value $\vec{\mu}$ of the source parameters is not known prior to the measurement. This operation must therefore be repeated for a large number of trial values $\vec{\mu}_{\rm trial}$; the corresponding signals $h(t;\vec{\mu}_{\rm trial})$ are known as {\it templates}. The overlap integral defines the function $S(\vec{\mu}_{\rm trial})$ given by \begin{eqnarray} S(\vec{\mu}_{\rm trial}) &=& \int h(t;\vec{\mu}_{\rm trial}) w(t-t') s(t')\, dt'\, dt \nonumber \\ &=& \bigl\langle h(\vec{\mu}_{\rm trial}) \bigm| s \bigr\rangle. \label{15} \end{eqnarray} To obtain the second line we have inserted the Fourier decompositions of $h$, $w$, and $s$, and carried out the integrations over time. The inner product $\langle \cdot | \cdot \rangle$ is defined as \begin{equation} \langle a | b \rangle = 2 \int_0^\infty \frac{\tilde{a}^*(f) \tilde{b}(f) + \tilde{a}(f) \tilde{b}^*(f)}{S_n(f)}\, df, \label{16} \end{equation} where $a$ and $b$ are arbitrary functions of time, with Fourier transforms $\tilde{a}$ and $\tilde{b}$. A quantity analogous to $S(\vec{\mu}_{\rm trial})$ can be defined for the noise alone, by replacing $s$ to the right of Eq.~(\ref{15}) by $n$. Operationally, this amounts to filtering the detector output when a gravitational-wave signal is known {\it not} to be present. Because the noise is a random function, the integrals $\int h(t;\vec{\mu}_{\rm trial}) w(t-t') n(t')\, dt'\, dt$ are random also. And because the noise is assumed to have zero mean, to take an average over all possible realizations of the noise (by repeating the measurements many times) would yield a zero value. None of these quantities would be especially useful. We are therefore led to consider the root-mean-square average of these integrals, \begin{eqnarray} N(\vec{\mu}_{\rm trial}) &=& \mbox{rms} \int h(t;\vec{\mu}_{\rm trial}) w(t-t') n(t')\, dt'\, dt \\ \nonumber &=& \sqrt{ \bigl\langle h(\vec{\mu}_{\rm trial}) \bigm| h(\vec{\mu}_{\rm trial}) \bigr\rangle } \equiv \rho(\vec{\mu}_{\rm trial}). \label{17} \end{eqnarray} To go from the first to the second line requires some machinery which will not be presented in this review. We refer the reader to Ref.~\cite{Wainstein} for the missing steps. \section{The signal-to-noise ratio} The {\it signal-to-noise ratio} is defined to be the ratio of $S$ over $N$: \begin{equation} \mbox{\small SNR}(\vec{\mu}_{\rm trial}) = \frac{ \bigl\langle h(\vec{\mu}_{\rm trial}) \bigm| s \bigr\rangle}{\rho(\vec{\mu}_{\rm trial})}. \label{18} \end{equation} It is clear that $\mbox{\small SNR}(\vec{\mu}_{\rm trial})$ is a random function, since $s(t)$ is itself a random function, being the superposition of signal plus noise. Its expectation value, or average over all possible realizations of the noise, is zero in the absence of signal (since $\overline{n} = 0$), and \begin{equation} \overline{\mbox{\small SNR} (\vec{\mu}_{\rm trial})} = \frac{ \bigl\langle h(\vec{\mu}_{\rm trial}) \bigm| h(\vec{\mu}) \bigr\rangle}{ \rho(\vec{\mu}_{\rm trial})} \label{19} \end{equation} in the presence of the signal $h(t;\vec{\mu})$. It is important to notice that in Eq.~(\ref{19}), the numerator is the overlap integral between the true signal $h(t;\vec{\mu})$ and the templates $h(t;\vec{\mu}_{\rm trial})$. In the absence or presence of a signal, the variance in the signal-to-noise ratio is precisely equal to unity \cite{CutlerFlanagan}, independently of the values of $\vec{\mu}$ and $\vec{\mu}_{\rm trial}$. This fully summarizes the statistical properties of the signal-to-noise ratio, since $\mbox{\small SNR}(\vec{\mu}_{\rm trial})$ is a Gaussian random function. [This can be seen from the fact that $s(t)$ itself is Gaussian, being the superposition of signal plus Gaussian noise.] It is intuitively clear that choosing a template with $\vec{\mu}_{\rm trial} = \vec{\mu}$ will produce the largest possible expectation value of the signal-to-noise ratio. This statement, known as {\it Wiener's theorem} \cite{Wainstein}, can easily be shown to be true by applying Schwarz's inequality to the right-hand side of Eq.~(\ref{19}). We have \begin{equation} \mbox{max} \Bigl\{ \overline{ \mbox{\small SNR} (\vec{\mu}_{\rm trial})} \Bigr\} = \overline{\mbox{\small SNR} (\vec{\mu})} = \rho(\vec{\mu}) = \sqrt{ \bigl\langle h(\vec{\mu}) \bigm| h(\vec{\mu}) \bigr\rangle }. \label{20} \end{equation} The fact that the signal-to-noise ratio has a variance of unity indicates that a signal can be concluded to be present only if $\rho(\vec{\mu})$ is significantly larger than 1. We will come back to this point in the next section. The true value of the source parameters can therefore be determined by maximizing $\mbox{\small SNR} (\vec{\mu}_{\rm trial})$ over all possible values of the trial parameters. The fact that the signal-to-noise ratio has a variance of unity implies that this determination can only have a limited degree of accuracy. The statistical errors decrease with increasing $\rho(\vec{\mu})$; since this is proportional to the signal's amplitude, a stronger signal gives better accuracy. Maximizing the signal-to-noise ratio is essentially equivalent to maximizing the overlap integral \[ \bigl\langle h(\vec{\mu}_{\rm trial}) \bigm| h(\vec{\mu}) \bigr\rangle. \] It is easy to see how the choice of the parameters $\vec{\mu}_{\rm trial}$ can affect the overlap integral. Consider a toy waveform with three parameters: arrival time, initial phase, and chirp mass. A mismatch in the arrival times clearly reduces the overlap integral: the signal and the template, taken to be functions of time, might have support in entirely different regions of the time axis, leading to a vanishing overlap. Supposing that the arrival times are matched, a mismatch in the initial phases can also reduce the overlap integral: the signal and the template might be out of phase with each other, leading to an oscillating integrand and a vanishingly small overlap. Finally, supposing that both the arrival times and initial phases are matched, a mismatch in the chirp masses would also reduce the overlap integral. This is because the chirp mass governs the rate at which the signal's frequency changes with time; cf.~Eq.~(\ref{9}). Signal and template, starting at the same time with the same phase, might thereafter go out of phase, thereby reducing the overlap. \section{Signal detection} The first order of business when analysing the output of a gravitational-wave detector is to decide whether or not a signal is present. Here we assume that the signal must be of a specific form, corresponding to a coalescing binary system. In this section we discuss signal detection --- how the technique of matched filtering can be employed to find the signal in the noisy data stream. In the next section we will discuss signal measurement --- how matched filtering is used to estimate the value of the source parameters once the signal has been found. As mentioned in the previous section, a signal can be concluded to be present if the maximum value of the signal-to-noise ratio, $\mbox{\small SNR} \equiv \mbox{max}\{\mbox{\small SNR}(\vec{\mu}_{\rm trial})\}$, is significantly larger than unity. In fact, there exists a threshold value $\mbox{\small SNR}^*$ such that a signal is concluded to present, with a certain confidence level, if $\mbox{\small SNR} > \mbox{\small SNR}^*$. To figure out how large this threshold must be is a standard application of the statistical theory of signal detection \cite{Helstrom}, which was developed largely for the purpose of detecting radar signals. The theory can easily be taken over to the case gravitational-wave signals \cite{Finn}. To go into the detail of this theory would be outside the scope of this review. We shall simply state that $\mbox{\small SNR}^*$ is fixed by selecting a small, acceptable value for the probability that a signal would falsely be concluded to be present --- the false alarm probability. (This is the Neyman-Pearson criterion \cite{Helstrom}, which is more precisely formulated in terms of the likelihood ratio, the ratio of the probability that a signal is present to the probability that it is absent.) Once $\mbox{\small SNR}^*$ is fixed, the level of confidence that a signal is indeed present increases with $\mbox{\small SNR} > \mbox{\small SNR}^*$. A typical ballpark value for the threshold is $\mbox{\small SNR}^* = 6$. In the preceding paragraph it was assumed that the signal-to-noise ratio $\mbox{\small SNR}$ is computed using template waveforms which are functions of the parameters $\vec{\mu}$ introduced in Sec.~7. These parameters have direct physical meaning; they include the chirp mass and other meaningful parameters to be introduced in Sec.~12. However, since these parameters are {\it not} estimated during the detection stage of the data analysis (they are estimated only {\it after} a signal has been found), there is no particular need to parametrize the templates with $\vec{\mu}$. In fact, it may be desirable, in order to minimize the computational effort, to parametrize the signal in a completely different way. The new parameters, $\vec{\alpha}$, would then have no particular physical significance. What is required is that the new templates, $h(t;\vec{\alpha})$, reproduce the behaviour of the expected gravitational-wave signal. In other words, the templates $h(t;\vec{\mu})$ and $h(t;\vec{\alpha})$ should span the same ``signal space'', but $h(t;\vec{\alpha})$ should do so most efficiently. Unlike $h(t;\vec{\mu})$, $h(t;\vec{\alpha})$ need not be derived from the field equations of general relativity. With these new {\it detection templates}, the signal-to-noise ratio is defined as \begin{equation} \mbox{\small SNR} = \mbox{max} \bigl\{ \mbox{\small SNR} (\vec{\alpha}) \bigr\}, \label{21} \end{equation} where $\mbox{\small SNR}(\vec{\alpha})$ is defined as in Eq.~(\ref{18}). As before, a signal is concluded to be present if $\mbox{\small SNR} > \mbox{\small SNR}^*$. \section{Signal measurement} Once the detection templates $h(t;\vec{\alpha})$ have been used to conclude that a signal is present in the data stream, they are replaced with the {\it measurement templates} $h(t;\vec{\mu}_{\rm trial})$ in order to estimate the value of the physical parameters $\vec{\mu}$. The procedure to estimate the source parameters was explained in Sec.~8. As was mentioned, the idea is to maximize over all possible values of the trial parameters the overlap integral \[ \bigl\langle h(\vec{\mu}) \bigm| h(\vec{\mu}_{\rm trial}) \bigr\rangle . \] We already have discussed the effect of a mismatch in the value of the parameters. What remains to be discussed is the effect of a mismatch in the functional form of the template with respect to that of the true waveform. The true waveform is governed by the exact laws of general relativity. The template, on the other hand, is necessarily constructed using an approximation to the exact laws. (That approximations must be made in analytic calculations is obvious; in a numerical treatment the approximation resides in the finite differencing of the field equations.) This, clearly, must have an effect on our strategy for extracting the information contained in the gravitational-wave signal. This can be seen simply from the fact that Wiener's theorem, as stated in Sec.~8, strictly requires signal and template to have the same functional form; they are allowed to differ only in the value of their parameters. Suppose that gravitational waves coming from a given source are received without noise, so that the true signal $h(t;\vec{\mu})$ is measured accurately. Suppose also that the true value $\vec{\mu}$ of the source parameters is known (God has spoken). Then a computation of the overlap integral with $\vec{\mu}_{\rm trial} = \vec{\mu}$ but with an approximate template $h(t;\vec{\mu}_{\rm trial})$ will not yield the maximum possible value for $\mbox{\small SNR}(\vec{\mu}_{\rm trial})$. This is because the approximation differs from the true signal, both in amplitude and in phase. Since both signal and template undergo a large number of oscillations (recall that for a system of two neutron stars, this number is approximately 16 000), the overlap integral is most sensitive to phase differences: a slight phase lag causes the integrand to oscillate, thereby severely reducing the signal-to-noise ratio with respect to its maximum possible value. A gravitational-wave astronomer doesn't know before the measurement the true value of the source parameters, and must work with an approximation to the true general-relativistic waveform. We have seen that the phase lag occurring between the true signal and the approximate template when the parameters are matched reduces the signal-to-noise ratio from its maximum possible value. It follows that maximizing $\mbox{\small SNR}(\vec{\mu}_{\rm trial})$ with approximate templates introduce {\it systematic errors} into the estimation of the source parameters: evaluating the signal-to-noise ratio with $\vec{\mu}_{\rm trial} = \vec{\mu} + \delta \vec{\mu}$ will return, for some $\delta \vec{\mu}$, a number larger than $\mbox{\small SNR}(\vec{\mu})$. The systematic errors are precisely the value of $\delta \vec{\mu}$ for which the signal-to-noise ratio is largest. If the templates are a poor approximation to the true signal, then the systematic errors will be larger than the statistical errors arising because $\mbox{\small SNR}(\vec{\mu}_{\rm trial})$ is a random function (see Sec.~8). We therefore appreciate the need for constructing measurement templates which are as accurate as possible, especially in phase \cite{FinnChernoff,Cutleretal}. The requirement is that the systematic errors in the estimated parameters must be smaller than the statistical errors. An estimate for the required degree of accuracy comes from the observation that the overlap integral will be significantly reduced if the template loses phase by as much as one wave cycle with respect to the true signal. Since the total number of wave cycles is approximately 16 000, we have \begin{equation} \mbox{accuracy} \sim \frac{1}{16\ 000} \sim 10^{-4}. \label{22} \end{equation} Since the orbital velocity $v$ is of order $10^{-1}$ when the gravitational-wave frequency is in the relevant bandwidth, relativistic corrections {\it at least} of order $v^4$ are required to improve the quadrupole-formula expression given in Sec.~6. As we shall see, this is an underestimate. \section{Waveform calculations: post-Newtonian theory} We have seen that accurate measurement templates are required to make the most of the gravitational-wave signals originating from coalescing compact binaries. We also have seen that the quadrupole-formula waveform, Eq.~(\ref{7}), is not sufficiently accurate; relativistic corrections at least of order $v^4$ are required. How does one go about calculating these? A possible line of approach is to use a slow-motion approximation to the equations of general relativity. This is based on the requirement that if $v$ is a typical velocity inside the matter source, then \begin{equation} v \ll 1. \label{23} \end{equation} We shall call this approximation ``post-Newtonian theory'' \cite{Will}. We point out that for binary systems, post-Newtonian theory makes no assumption regarding the relative size of the two masses. This is to be contrasted with the perturbation approach, discussed in Sec.~13, in which the mass ratio is assumed to be small, but no restriction is put on $v$. In post-Newtonian theory \cite{BlanchetDamour}, one starts by defining fields $h^{\alpha\beta}$ as \begin{equation} h^{\alpha\beta} = \sqrt{-g}\, g^{\alpha\beta} - \eta^{\alpha\beta}, \label{24} \end{equation} where $g^{\alpha\beta}$ is the inverse of the true metric $g_{\alpha\beta}$ with determinant $g$, and $\eta^{\alpha\beta}$ is the metric of Minkowski spacetime. When the harmonic gauge conditions \begin{equation} \partial_\beta h^{\alpha\beta} = 0 \label{25} \end{equation} are imposed, the {\it exact} Einstein field equations reduce to \begin{equation} \Box h^{\alpha\beta} = 16\pi (-g) T^{\alpha\beta} + \Lambda^{\alpha\beta}. \label{26} \end{equation} Here, $\Box=\eta^{\alpha\beta} \partial_\alpha \partial_\beta$ is the flat-spacetime wave operator, $T^{\alpha\beta}$ is the stress-energy tensor of the source, and $\Lambda^{\alpha\beta}$ is nonlinear in $h^{\alpha\beta}$ and represents an effective stress-energy tensor for the gravitational field. The detailed way in which one solves these equations is quite complicated, and will not be described here. The essential ideas are these \cite{BlanchetDamour}: One first integrates the equations in the near zone ($r < \lambda$, where $r$ is the flat-space radius and $\lambda$ the gravitational wavelength) assuming slow motion, or $\partial h^{\alpha\beta} / \partial t \ll \partial h^{\alpha\beta} / \partial x^i$. One does this by iterations: the nonlinear terms in Eq.~(\ref{26}) are first neglected, and the resulting linear equations integrated. These solutions are then used as input for the next iteration. This process is continued until the desired degree of accuracy is obtained. This is the standard post-Newtonian approach \cite{Damour}. One next integrates the equations everywhere in the vacuum region outside the source. This is done once again by iterations, assuming $h^{\alpha\beta} \ll 1$, but assuming nothing about the relative size of $\partial h^{\alpha\beta} / \partial t$ with respect to the spatial derivatives. This is because the vacuum region contains the wave zone, in which the field propagates with the speed of light; a slow-motion assumption would therefore not do for the field itself. This is the post-Minkowskian approach \cite{Damour}. Using the post-Minkowskian approach one constructs, by successive approximations, the most general solution to the Einstein equations outside the source. This is characterized by two infinite sets of arbitrary multipole moments \cite{Thorne80}, the mass moments $M_{\ell m}(t-r)$ and the current moments $J_{\ell m}(t-r)$, were $\ell$ and $m$ are the standard spherical-harmonic indices. (In practice, the fields $h^{\alpha\beta}$ are equivalently expressed in terms of symmetric-trace-free moments, not spherical-harmonic moments.) The general solution is then matched to the near-zone solution in the region of common validity, and the multipole moments are thus determined. Finally, one expresses the radiation field --- the time-varying, $O(1/r)$ part of the gravitational field --- in terms of the derivatives of the mass and current multipole moments \cite{Thorne80}. This gives the gravitational waveform. The gravitational-wave luminosity $dE/dt$ can also be obtained from the radiation field. \section{Waveform to second post-Newtonian order} To date, the post-Newtonian calculation of the waveform has been carried out accurately through second post-Newtonian order --- $O(v^4)$ --- beyond the leading-order, quadrupole-formula expressions given in Sec.~6. The complete waveform will not be displayed here. Instead, we will focus solely on the waveform's {\it phasing}. The phasing of the waves can be determined from $df/dt$, the rate of change of the gravitational-wave frequency. This can be expressed as \begin{equation} \frac{df}{dt} = \frac{dE/dt}{dE/df}, \label{27} \end{equation} where $dE/dt$ is (minus) the gravitational-wave luminosity, and $dE/df$ relates orbital energy to orbital frequency ($f$ is twice the orbital frequency; the orbits are assumed to be circular). Both these quantities can be expanded in powers of \begin{equation} v \equiv (\pi M f)^{1/3}, \label{28} \end{equation} with leading-order terms \cite{MTW} \begin{equation} \biggl( \frac{dE}{dt} \biggr)_{\!\!N} = -\frac{32}{5} \eta^2 v^{10}, \qquad \biggl( \frac{dE}{df} \biggr)_{\!\!N} = -\frac{\pi}{3} \mu M v^{-1}, \label{29} \end{equation} where the subscript $N$ stands for ``Newtonian''. We have introduced the reduced mass $\mu$, the total mass $M$, and the mass ratio $\eta$ as \begin{equation} \mu = \frac{m_1 m_2}{m_1 + m_2}, \qquad M = m_1 + m_2, \qquad \eta = \frac{\mu}{M}. \label{30} \end{equation} Notice that $\eta$ is restricted to the interval $0 < \eta \leq 1/4$, with $\eta=1/4$ for $m_1=m_2$. It is easy to check that Eqs.~(\ref{27}) and (\ref{29}) reproduce Eq.~(\ref{9}) above; the chirp mass can be expressed as ${\cal M} = \eta^{3/5} M$. For simplicity we will, in the following, focus on the quantity $dE/dt$. To second post-Newtonian order, the luminosity takes the form \begin{equation} \frac{dE}{dt} = \biggl( \frac{dE}{dt} \biggr)_{\!\!N} \Biggl[ 1 - \biggl( \frac{1247}{336} + \frac{35}{12}\, \eta \biggr) v^2 + \bigl( 4\pi - \mbox{\small SO} \bigr) v^3 - \biggl( \frac{44711}{9072} - \frac{9271}{504}\, \eta - \frac{65}{18}\, \eta^2 + \mbox{\small SS} \biggr) v^4 + \cdots \Biggr]. \label{31} \end{equation} Here, the terms {\small SO} and {\small SS} are due to spin-orbit and spin-spin interactions, respectively \cite{KWW}; these occur if the masses $m_1$ and $m_2$ are rotating. Let $\vec{S}_1$ and $\vec{S}_2$ be the spin angular momentum of each mass, and define the dimensionless quantities $\vec{\chi}_a = \vec{S}_a/{m_a}^2$, for $a=\{1,2\}$. Let also $\hat{L}$ be the direction of orbital angular momentum. Then \cite{KWW} \begin{equation} \mbox{\small SO} = \frac{1}{4} \sum_{a=1}^2 \Bigl[ 11 (m_a/M)^2 + 5\eta \Bigr] \hat{L} \cdot \vec{\chi}_a \label{32} \end{equation} is the spin-orbit term, and \begin{equation} \mbox{\small SS} = \frac{\eta}{48} \Bigl( 103 \vec{\chi_1} \cdot \vec{\chi_2} - 289 \hat{L} \cdot \vec{\chi}_1 \, \hat{L} \cdot \vec{\chi}_2 \Bigr) \label{33} \end{equation} is the spin-spin term. In Eq.~(\ref{31}), the leading-order term was first calculated in 1963 by Peters and Mathews \cite{PetersMathews} using the standard quadrupole formula. The first post-Newtonian correction, at order $v^2$, was calculated in 1976 by Wagoner and Will \cite{WagonerWill}. The $4\pi v^3$ term is due to wave propagation effects: As the waves propagate out of the near zone they are scattered by the curvature of spacetime, and this modifies both the amplitude and the phase of the waveform; this ``tail term'' was first calculated in 1993 by this author \cite{paperI}, and then independently by Wiseman \cite{Wiseman} and Blanchet and Sch\"afer \cite{BlanchetSchafer}. The second post-Newtonian correction, at order $v^4$, was calculated in 1994 by Blanchet, Damour, Iyer, Will, and Wiseman \cite{BDIWW}. Finally, the spin-orbit and spin-spin corrections were calculated in 1993 by Kidder, Will, and Wiseman \cite{KWW}. We see from Eq.~(\ref{31}) that the post-Newtonian corrections bring a number of additional source parameters into the picture. The waveform no longer depends uniquely upon the chirp mass ${\cal M}$; it now depends upon the masses $m_1$ and $m_2$ separately, and upon the spin-orbit and spin-spin parameters (which stay approximately constant as the system evolves toward coalescence). These new parameters must be included into $\vec{\mu}$ when the signal is analyzed using matched filtering. What we have at this point is an expression for the waveform which is accurate to second post-Newtonian order. The question facing us is whether this waveform is sufficiently accurate to be used as measurement templates. In other words, are the systematic errors generated by these templates guaranteed to be smaller than the statistical errors? Evidently, to answer this question is difficult, since we do not have access to the exact waveform in order to make comparisons. In the next section we will consider a model problem for which the waveform {\it can} be calculated exactly, thereby enabling us to judge the accuracy of the post-Newtonian expansion. We will find, in Sec.~15, that the answer to this question is, most likely, no: the second post-Newtonian waveform is not sufficiently accurate. \section{Waveform calculations: perturbation theory} A different line of attack for solving Einstein's equations for a compact binary system is to assume that one of the bodies is very much less massive than the other \cite{Poisson}. We therefore demand \begin{equation} \mu / M \ll 1, \label{34} \end{equation} where $\mu$ is the reduced mass and $M$ the total mass. In this limit $\mu$ is practically equal to the smaller mass $m_1$, and $M$ is practically equal to the larger mass $m_2$. In contrast with the post-Newtonian approach, we assume {\it nothing} about the size of the velocity $v$. This approach takes advantage of the fact that when Eq.~(\ref{34}) is valid, the smaller mass creates only a small perturbation in the gravitational field of the larger mass, which can be taken to be the Schwarzschild field. The total gravitational field can therefore be written as \begin{equation} g_{\alpha\beta} = g_{\alpha\beta}^{(0)} + h_{\alpha\beta}, \label{35} \end{equation} where $g_{\alpha\beta}^{(0)}$ represents the background Schwarzschild metric, and $h_{\alpha\beta}$ the perturbation. Writing Einstein's equations for $g_{\alpha\beta}$ and linearizing with respect to $h_{\alpha\beta}$, one finds that the perturbation must satisfy an inhomogeneous wave equation in the Schwarzschild spacetime. Schematically, \begin{equation} \Box^{\alpha\beta\mu\nu} h_{\mu\nu} = 8\pi T^{\alpha\beta}, \label{36} \end{equation} where $\Box^{\alpha\beta\mu\nu}$ is an appropriate curved-spacetime wave operator, and $T^{\alpha\beta}$ the stress-energy tensor associated with the orbiting mass. We will specifically assume that the central body is a Schwarzschild black hole of mass $M$. This assumption is made for simplicity, and removes the need to model the star's interior. As a matter of fact, the internal structure of the bodies is irrelevant, except during the last few orbital cycles before coalescence; this was discussed in Sec.~5. Taking advantage of this, we model the orbiting body as a point particle of mass $\mu$, so that its stress-energy tensor is a Dirac distribution with support on the particle's world line. For simplicity, and also because it is physically well motivated (as explained in Sec.~5), we take the world line to be a circular geodesic of the Schwarzschild spacetime. With $\{t,r,\theta,\phi\}$ as the usual Schwarzschild coordinates, $r_0$ denotes the orbital radius, and $\Omega = d\phi/dt$ is the angular velocity. We have \begin{equation} v \equiv \Omega r_0 = (M/r_0)^{1/2} = (M\Omega)^{1/3}, \label{37} \end{equation} and the gravitational-wave frequency $f$ is given by $2 \pi f= 2\Omega$. We stress once more that in the perturbation approach, $v$ is not required to be small. The only limitation on $v$ comes from the fact that for $r_0 \leq 6M$, circular orbits are no longer stable; this implies $v < 6^{-1/2} \simeq 0.4082$. Black-hole perturbations are conveniently treated with the Teukolsky formalism \cite{Teukolsky}, in which gravitational perturbations are represented by the complex-valued function $\Psi_4$, a particular component of the perturbed Weyl tensor; the tensor $h_{\alpha\beta}$ can be reconstructed from $\Psi_4$. The equation satisfied by $\Psi_4$ admits a separation of the variables. When $\Psi_4$ is expanded in spherical harmonics and decomposed into Fourier modes $e^{-i\omega t}$, one obtains an ordinary second-order differential equation --- the Teukolsky equation --- for the radial function $R_{\omega \ell m}(r)$. Here, $\ell$ and $m$ are the usual spherical-harmonic indices. Schematically, and omitting the subscript $\omega \ell m$, this equation takes the form \begin{equation} {\cal D} R(r) = T(r), \label{38} \end{equation} where $\cal D$ is a second-order differential operator, and $T(r)$ the source, constructed from the particle's stress-energy tensor \cite{Poisson}. Equation (\ref{38}) can be integrated in the standard way by constructing a Green's function $G(r,r')$ out of two linearly independent solutions to the homogeneous equation. These solutions, $R_<(r)$ and $R_>(r)$, respectively satisfy appropriate boundary conditions at the inner ($r=2M$) and outer ($r=\infty$) boundaries. Schematically, $G(r,r') = R_<(r_<) R_>(r_>)$, where $r_<$ ($r_>$) is the lesser (greater) of $r$ and $r'$. The solution to Eq.~(\ref{38}) can then be expressed as $R(r) = \int G(r,r') T(r')\, dr'$, and $\Psi_4$ can be reconstructed by summing over all the modes. Finally, the gravitational waveform $h$ and the luminosity $dE/dt$ can be obtained from the asymptotic behaviour of $\Psi_4$ when $r\to\infty$. To integrate Eq.~(\ref{38}) therefore reduces to solving the homogeneous Teukolsky equation, ${\cal D} R(r)=0$, for the functions $R_<(r)$ and $R_>(r)$. This, it turns out, is equivalent to integrating the Regge-Wheeler equation \cite{ReggeWheeler} \begin{equation} \Biggl\{ \frac{d^2}{dr^{*2}} + \omega^2 - \biggl(1 - \frac{2M}{r} \biggr) \biggl[ \frac{\ell(\ell+1)}{r^2} - \frac{6M}{r} \biggr] \Biggr\} X_{\omega\ell}(r) = 0, \label{39} \end{equation} for the functions $X_<(r)$ and $X_>(r)$; here, $d/dr^* = (1-2M/r) d/dr$. This comes about because a solution to the homogeneous Teukolsky equation can easily be related to a solution to the Regge-Wheeler equation. The relation is known as the Chandrasekhar transformation \cite{Chandra}. \section{Luminosity from the perturbation approach} The problem of calculating the gravitational waveform for a compact binary system with small mass ratio can therefore be reduced to the simple one of integrating Eq.~(\ref{39}). The relevant dimensionless parameter entering this equation is $M\omega$, where, for circular orbits, $\omega = m\Omega$ \cite{Poisson}. We therefore have, using Eq.~(\ref{37}), $M\omega = m v^3$. For arbitrary values of $M\omega$ the Regge-Wheeler equation must be integrated numerically \cite{paperII}. This must be done separately for each selected value of $v$ in the interval $0.1 < v < 0.4$ (approximately corresponding, for systems of a few solar masses, to the frequency interval $10\ \mbox{Hz} < f < 1\ 000\ \mbox{Hz}$). \cite{Reif} \begin{figure}[t] \special{hscale=70 vscale=70 hoffset=50.0 voffset=-500.0 angle=0.0 psfile=fig2.ps} \vspace*{3.2in} \caption[Fig. 2]{Various representations of $(dE/dt)/(dE/dt)_N$ as a function of orbital velocity $v$. The solid curve represents the exact results, obtained numerically. The various broken curves represent the various post-Newtonian approximations, as explained in the text.} \end{figure} Once again we will focus on $dE/dt$, the gravitational-wave luminosity. In Fig.~2 we present a plot of $(dE/dt)/(dE/dt)_N$ as a function of $v$. This is the ratio of the luminosity as calculated exactly (numerically) using the perturbation approach, to the quadrupole-formula expression \begin{equation} \biggl( \frac{dE}{dt} \biggl)_{\!\!N} = - \frac{32}{5} \Bigl(\frac{\mu}{M} \Bigr)^2\, v^{10}. \label{40} \end{equation} The numerical results are depicted as a solid curve. Apart from negligible numerical errors, these results are {\it exact} to all orders in $v$; the only assumption used in the calculation concerns the smallness of $\mu/M$. The various broken curves will be described below. The figure shows that $(dE/dt)/(dE/dt)_N$ tends toward unity as $v\to 0$, and stays within 15\% of unity everywhere in the interval $0 < v < 6^{-1/2}$. It is now easy to judge the accuracy of the post-Newtonian expansion for $dE/dt$ in the limit $\mu/M \to 0$. One simply evaluates Eq.~(\ref{31}) in this limit, with $\mbox{\small SO} = \mbox{\small SS} = 0$, and compare with the numerical results. This post-Newtonian curve is labeled ``4'' in the figure, and we see that the numerical results are only imperfectly reproduced. The consequences of this will be discussed in the next section. One can in fact do better than this. The perturbation approach is not only suitable for exact, numerical computations. By combining it with a slow-motion approximation --- putting $v \ll 1$ on top of $\mu/M \ll 1$ --- it also becomes suitable for approximate, analytical computations. Methods for integrating the Regge-Wheeler equations analytically in the limit $M\omega \ll 1$ were devised by various authors \cite{paperI,Poisson,Sasaki}. The basic idea is to proceed by iterations: Eq.~(\ref{39}) with $M\omega=0$ is solved in terms of spherical Bessel functions, and this zeroth order solution is used as input for the first iteration. Using these methods, Tagoshi and Sasaki \cite{TagoshiSasaki} were able to calculate $dE/dt$ analytically, accurately through fourth post-Newtonian order, two full orders beyond Eq.~(\ref{31}). They derive the rather impressive expression \begin{eqnarray} \frac{dE}{dt} &=& \biggl( \frac{dE}{dt} \biggr)_{\!\!N} \Biggl[ 1 - \frac{1247}{336}\, v^2 + 4\pi\, v^3 - \frac{44711}{9072}\, v^4 - \frac{8191}{672}\, \pi\, v^5 \nonumber \\ & & \mbox{} + \biggl( \frac{6643739519}{69854400} - \frac{1712}{105}\, \gamma + \frac{16}{3}\, \pi^2 - \frac{3424}{105}\, \ln 2 - \frac{1712}{105}\, \ln v \biggr)\, v^6 - \frac{16285}{504}\, \pi \, v^7 \nonumber \\ & & \mbox{} + \biggl( - \frac{323105549467}{3178375200} + \frac{232597}{4410}\, \gamma - \frac{1369}{126}\, \pi^2 + \frac{39931}{294}\, \ln 2 - \frac{47385}{1568}\, \ln 3 + \frac{232597}{4410}\, \ln v \biggr)\, v^8 \nonumber \\ & & \mbox{} + \cdots \Biggr]. \label{41} \end{eqnarray} Notice the presence of $\ln v$ terms in this expansion, as well as that of the Euler number $\gamma \simeq 0.5772$. Notice also that the first four terms reproduce Eq.~(\ref{31}) in the limit $\mu/M \to 0$, when $\mbox{\small SO} = \mbox{\small SS} = 0$. This, of course, is as it should be. The broken curves in Fig.~2 represent plots of Eq.~(\ref{41}) truncated to various orders in $v$. For example, the curve labeled ``6'' is a plot of Eq.~(\ref{41}) with all terms of order $v^7$ and $v^8$ discarded. We see from the figure that the post-Newtonian expansion converges poorly. (The suspicion, in fact, is that the series is only asymptotic.) Witness in particular the poor quality of the curve ``5'' with respect to ``4'', and compare also ``8'' to ``7''. \section{Accuracy of the post-Newtonian expansion} We have seen that in the $\mu/M \to 0$ limit, the post-Newtonian expansion for $dE/dt$ converges poorly. A similar statement can be made about $dE/df$: the post-Newtonian expansion \begin{equation} \frac{dE}{df} = \biggr( \frac{dE}{df} \biggr)_{\!\!N} \biggl( 1 - \frac{3}{2}\, v^2 - \frac{81}{8}\, v^4 - \frac{675}{16}\, v^6 - \frac{19845}{128}\, v^8 + \cdots \biggr), \label{42} \end{equation} where \begin{equation} \biggl( \frac{dE}{df} \biggr)_{\!\!N} = -\frac{\pi}{3} \mu M v^{-1}, \label{43} \end{equation} converges slowly to the exact result \cite{foot2} \begin{equation} \frac{dE}{df} = \biggr( \frac{dE}{df} \biggr)_{\!\!N} \bigl( 1 - 6v^2 \bigr) \bigl( 1 - 3v^2 \bigr)^{-3/2}. \label{44} \end{equation} Contrary to Eq.~(\ref{41}), the expansion (\ref{42}) is actually {\it known} to converge (for all values of $v$ in the interval $0 < v < 6^{-1/2}$). In this section we address the issue as to how much of an obstacle the poor convergence of the post-Newtonian expansion poses to the construction of accurate measurement templates \cite{paperVI}. We will answer this question for binary systems with small mass ratios, using the results described in the preceding section. Since there is no reason to believe that the convergence of the post-Newtonian expansion would be much improved for systems of comparable masses, our results should also apply, at least qualitatively, for such systems. One way to address this question is to ask, given a waveform constructed from the exact numerical results, how much signal-to-noise ratio is lost by matched filtering the exact signal with an approximate post-Newtonian template? {}From Sec.~8 we know that filtering with the exact signal $h(t)$ would give $\mbox{\small SNR}|_{\rm max}$, the largest possible value of the signal-to-noise ratio. On the other hand, filtering with a post-Newtonian template $h_n$, where $n$ denotes the order in $v$ to which the expansion is taken (for example, $n=4$ represents a waveform accurate to second post-Newtonian order), gives the smaller value $\mbox{\small SNR}|_{\rm actual}$. From the results of Sec.~7 and 8 we obtain \begin{equation} {\cal R}_n \equiv \frac{ \mbox{\small SNR}|_{\rm actual} }{ \mbox{\small SNR}|_{\rm max} } = \frac{ | \langle h | h_n \rangle |}{ \sqrt{ \langle h | h \rangle \langle h_n | h_n \rangle }}. \label{45} \end{equation} The detail of how to compute ${\cal R}_n$ is presented in Ref.~\cite{paperVI}. The calculation described here can only be carried out for binary systems with small mass ratios, because only for these do we have access to the exact waveform $h(t)$. Nevertheless, in the following we shall let $\mu/M$ become large, without altering our expressions for the exact and post-Newtonian waveforms. This is done without justification, but reflects the adopted point of view that the quality of the post-Newtonian approximation should not be appreciably affected by the finite-mass corrections. In doing this extrapolation, some thought must be given as to the interpretation of the ratio $\mu/M$. In the limit $\mu/M \to 0$ this can be taken to be both the ratio of the individual masses or the ratio of reduced mass to total mass. In the case of comparable masses, some choice must be made. We note that as $\mu/M$ is allowed to grow large, expressions (\ref{40}) and (\ref{43}) for the Newtonian quantities $(dE/dt)_N$ and $(dE/df)_N$ must be replaced by expressions (\ref{29}), {\it which are formally identical}. This shows that when extrapolating to the case of comparable masses, $\mu$ is to be interpreted as the {\it reduced mass}, and $M$ as the {\it total mass}. \begin{table}[t] \begin{center} \begin{tabular}{ccc} \hline \hline $n$ & PN & EXACT \\ \hline 4 & 0.5796 & 0.4958 \\ 5 & 0.4646 & 0.5286 \\ 6 & 0.7553 & 0.9454 \\ 7 & 0.7651 & 0.9864 \\ 8 & 0.7568 & 0.9695 \\ \hline \hline \end{tabular} \end{center} \caption[Table I]{Reduction in signal-to-noise ratio incurred when matched filtering with approximate, post-Newtonian templates. The first column lists the order $n$ of the approximation, the second column lists ${\cal R}_n$ as calculated using the post-Newtonian approximation for $dE/df$, and the third column lists ${\cal R}_n$ as calculated using the exact expression for $dE/df$.} \end{table} We quote the results for ${\cal R}_n$ corresponding to a system of two neutron stars, each of 1.4 $M_\odot$; these are displayed in the second column of Table 1. We see that even at quite a high order in the post-Newtonian expansion, only three quarters of the signal-to-noise ratio is reproduced by the post-Newtonian template. This shows that the apparently small discrepancies between the exact and post-Newtonian results for $dE/dt$ and $dE/df$ provide a serious obstacle to the construction of accurate measurement templates. It is interesting to ask how much of the signal-to-noise ratio would be recovered if $dE/df$ were kept exact instead of being expressed as a post-Newtonian expansion. The third column of the table displays the values of ${\cal R}_n$ calculated in this way. We see that most of the signal-to-noise is recovered: ${\cal R}_n$ can now be as large as 0.9864 instead of 0.7651. Why does the exact expression for $dE/df$ give such better results? It can be established \cite{paperVI} that this has to do with the following fact: While the exact expression for $dE/df$ correctly goes to zero at $v = 6^{-1/2}$ (at the innermost circular orbit), its post-Newtonian analogue fails to do so. [For example, the right-hand side of Eq.~(\ref{42}), truncated to order $v^8$, goes to zero at $v\simeq 0.4236 > 6^{-1/2}$, corresponding to a radius $r_0 \simeq 5.572 M$.] When corrected for this, the post-Newtonian template gives much better results. \section{Conclusion} We therefore see that the poor convergence of the post-Newtonian expansion is a serious obstacle to the construction of accurate measurement templates. Devising ways to extract the information contained in gravitational waves produced during the late inspiral of a compact binary system poses a great challenge to theoretical physicists. It is not clear that ``simply'' pushing to higher order in post-Newtonian theory will be enough to produce sufficiently accurate measurement templates (such that the systematic errors will be smaller than the statistical errors). There may be a need for theorists to develop alternative ways of dealing with this problem. It should be stressed that the convergence problem does not arise when constructing {\it detection templates}. Indeed, as was discussed in Sec.~9, there is no particular need for these templates to be derived from the equations of general relativity. And since the detection templates need not involve any parameters of direct physical significance, the notion of systematic errors does not apply to them. The only requirement for constructing detection templates is that they should span the appropriate ``signal space'', and that they should do so the most efficiently. The poor convergence of the post-Newtonian expansion is therefore not an obstacle for {\it detecting} gravitational waves from coalescing compact binaries. It is only an obstacle for extracting the information that the waves contain. To overcome this obstacle will undoubtedly be a theorist's challenge for years to come. This state of affairs is highly interesting from a historical point of view: Never before in the history of gravitational physics has experiment demanded such a high degree of sophistication on theoretical calculations. \section*{Acknowledgments} This work was supported by the Natural Sciences Foundation under Grant No.~PHY 92-22902 and the National Aeronautics and Space Administration under Grant No.~NAGW 3874. This article was completed while visiting the Theoretical Physics Institute of the University of Alberta; the author is most grateful to Werner Israel for his warm hospitality.
\section{Introduction} Most first-principles calculations on condensed matter are nowadays based on density functional theory (DFT)~\cite{hoh64,koh65,jon89,gil91,pay92}. This theory is formally exact, but in practice an approximation has to be made to the exchange-correlation energy, and the vast majority of calculations employ the local density approximation (LDA). The basic assumption is that the exchange-correlation energy per electron at any point in the system is related to the electron density at that point in the same way as in a uniform electron gas, and density gradients are ignored. Some theoretical justification can be given for this~\cite{har74,gun76,lan77}, and in practice the LDA works well in a wide range of situations. However, its accuracy is not always satisfactory, particularly when energy differences associated with changes of bonding are needed, as in e.g. molecular dissociation or the adsorption of molecules at surfaces. Attempts to improve the situation by adding lowest-order corrections in powers of the density gradient are not successful, but important progress has been made recently by requiring that the dependence of the energy on the gradients satisfies certain physical requirements. This has led to various forms of generalized gradient corrections (GGC)~\cite{lan83,per86,pw86,bec88,lac93}. In the last few years, there has been a large amount of work on the influence of different GGC schemes on the total energies of atoms and molecules~\cite{lan83,per86,pw86,lan85,kut88,bos90,mly91}, the equilibrium structure and cohesive energies of covalent crystals~\cite{gar92,kon90,ori92,jua93,sei95}, the ground state of iron~\cite{zhu92,hag93}, and the energetics of molecular adsorption on metal surfaces~\cite{whi94a,hu94,gun94}. However, so far as we are aware, there has been little work on the effect of GGC on the properties of partially ionic materials such as the oxides TiO$_2$ and SnO$_2$ treated here. The surface properties of materials like these are extremely important, because of their application as gas sensors and catalysts. We have recently reported a detailed study of the bulk and surface properties of SnO$_2$~\cite{IanYY,IanXX}, and we have initiated work on the interaction of molecules with the surfaces of both TiO$_2$ and SnO$_2$~\cite{unp}. An understanding of GGC is of considerable importance in this general area. The goal of the present paper is to study the effect on the bulk and surface properties of TiO$_2$ and SnO$_2$ of the two widely used GGC schemes due to Perdew and Wang~\cite{per86,pw86} and Becke and Perdew~\cite{per86,bec88}. \section{Techniques} The calculations are performed using the pseudopotential method~\cite{gil91,pay92}, so that only the valence electrons are represented explicitly, the valence-core interaction being represented by non-local norm-conserving pseudopotentials, which are generated by first-principles calculations on isolated atoms. Periodic boundary conditions are used, with the occupied electronic orbitals expanded in a plane-wave basis. The expansion includes all plane waves whose kinetic energy $\hbar^2 k^2 / 2m$ ($k$ the wavevector, $m$ the electronic mass) is less than a chosen cutoff energy $E_{\rm cut}$. The inclusions of gradient corrections within the pseudopotential plane-wave technique has recently been discussed in detail by White and Bird~\cite{whi94}, who show that a robust and accurate calculation of the GGC exchange-correlation energy and potential can be achieved by summation on exactly the same real-space grid as would be used for the LDA. This technique has been used in the present work. The first-principles pseudopotentials in Kleinman-Bylander representation~\cite{kle82} were generated using the optimization scheme of Lin {\em et al.}~\cite{lin93} in order to reduce the required value of the plane-wave cutoff $E_{\rm cut}$. The pseudopotentials used in the GGC calculations were constructed consistently by including gradient corrections in the generation scheme. The Sn pseudopotential was generated using the $5s^2 5p^2$ configuration for $s$- and $p$-wave components, and the $5s^15p^{0.5}5d^{0.5}$ configuration for the $d$-wave. The core radii were equal to 2.1, 2.1 and 2.5~a.u. for the $s$, $p$ and $d$ components respectively. The Ti pseudopotential was generated using the $4s^{1.85}3d^2$ configuration for $s$ and $d$ waves and the $4s^14p^{0.5}3d^{0.5}$ configuration for the $p$ wave, with core radii of 2.2, 1.5 and 2.4 a.u. for $s$, $p$ and $d$ waves respectively. The oxygen pseudopotential used in our LDA calculations was generated using the $2s^2 2p^4$ configuration for the $s$ and $p$ waves and the $2s^2 2p^{2.5} 3d^{0.5}$ configuration for the $d$ wave, with a single core radius of 1.65~a.u. For the gradient-corrected oxygen pseudopotential, we have used the single configuration $2s^2 2p^{3.5} 3d^{0.45}$ and the same core radius. The use of a core radius of 1.65~a.u. means that there is an appreciable overlap of the oxygen and metal core spheres in the SnO$_2$ and TiO$_2$ crystals, and in principle this could cause inaccuracies. However, direct comparisons of the present results with our earlier work on SnO$_2$~\cite{IanXX}, which employed an oxygen pseudopotential with the smaller core radius of 1.25~a.u., show that any errors due to core overlap are very small. The calculations have been done using a plane wave cut-off $E_{\rm cut}$ of 600~eV for SnO$_2$ and 1000~eV for TiO$_2$. Our tests show that with these cut-offs the energy per unit cell is converged to within 0.2~eV, the convergence with respect to $E_{\rm cut}$ being not noticeably influenced by the inclusion of gradient corrections, even though the gradient corrected pseudopotentials are less smooth and regular than the LDA ones~\cite{gar92,ort91}. The calculations were performed using the CETEP code~\cite{cla92} (the parallel version of the serial CASTEP code~\cite{pay92}) running on the 64-node Intel iPSC/860 machine at Daresbury Laboratory. The code uses the band-by-band conjugate-gradient technique to minimize the total energy with respect to plane-wave coefficients. The LDA calculations were performed using the Ceperley-Alder (CA) exchange-correlation function~\cite{cep80}. For the ground states calculations Brillouin zone sampling is performed using the lowest order Monkhorst-Pack set of k--points~\cite{mon76}, as in our earlier work on SnO$_2$~\cite{IanYY}. Electronic densities of states (DOS) associated with the ground state were calculated using the tetrahedron method~\cite{jep71,leh72}, with $k$-point sampling corresponding to 750 tetrahedra in the whole Brillouin zone. \section{Results and discussion} \subsection{Perfect SnO$_2$ and TiO$_2$ crystals} The 6-atom rutile unit cell of SnO$_2$ and TiO$_2$ is characterized by the two lattice parameters $a$ and $c$ and the internal parameter $u$: the positions of the four oxygens are $(\pm u, \pm u, 0)$, $( \frac{1}{2} \pm u, \frac{1}{2} \mp u, \frac{1}{2} )$. The equilibrium structure has then been determined by relaxation with respect to the lattice parameters $a$ and $c$ and the internal parameter $u$. The equilibrium values of these parameters both with and without gradient corrections are given in Table~\ref{tab1}. As usually happens, there is a tendency for the LDA to underestimate the lattice parameter. This is especially noticeable for SnO$_2$, where there may also be an effect due to our treatment of the 4$d$ shell as part of the core. The inclusion of gradient corrections tends to increase the lattice parameters, as has already been found for semiconducting and metallic systems~\cite{gar92,jua93,sei95}. The increase is 4~\% or more for the Perdew-Wang GGC, and leads to results for $a$ and $c$ that are appreciably greater than experimental values. For the Becke-Perdew GGC, the increase is roughly 3~\%. Both the $c/a$ ratio and the $u$ parameter are almost unaffected, and this suggests that the gradient corrections have the effect of an isotropic negative pressure, as pointed out by Seifert {\em et al.}~\cite{sei95}. We have calculated the electronic DOS for the SnO$_2$ perfect crystal using both LDA and the two GGC schemes, but the changes caused by GGC are very small and we do not show the results here. \subsection{The SnO$_2$ and TiO$_2$~(110) surfaces} Our calculations on the stoichiometric (110) surface of both materials have been done with the usual repeating slab geometry. The rutile structure can be regarded as consisting of (110) planes of atoms containing both metal (M) and oxygen (O) atoms, separated by planes containing oxygen alone, so that the sequence of planes is O - M$_2$O$_2$ - O - O - M$_2$O$_2$ - O etc. The entire crystal can then be built up of symmetrical 3-plane O - M$_2$O$_2$ - O units. The slabs we use contain three of these units, and our repeating cell contains 18 atoms (6~M and 12~O). The perfect (110) surface consists of rows of bridging oxygens lying above a metal-oxygen layer. The vacuum separating the slabs has been taken wide enough to ensure that interactions between neighboring slabs are small. The width we use corresponds to two O - M$_2$O$_2$ - O units, and is such that planes of bridging oxygens on the surfaces facing each other across the vacuum are separated by about 6.8~\AA. The surface structure has been determined by relaxing the entire system to equilibrium, and the calculations have been done with and without gradient corrections. As in our previous work on SnO$_2$~(110)~\cite{IanXX}, and the work of Ramamoorthy {\em et al.} on TiO$_2$~(110)~\cite{ram94a}, we find displacements of the surface atoms of order 0.1~\AA, with 5-fold and 6-fold coordinated metal atoms (M$_{\rm II}$ and M$_{\rm I}$) moving respectively into and out of the surface, in-plane oxygens (O$_{\rm II}$) moving out and bridging oxygens (O$_{\rm I}$) moving very little. The changes of the bond lengths between the surface atoms, including sub-bridging oxygens (O$_{\rm III}$) and the uppermost oxygens (O$_{\rm IV}$) of the following O - M$_2$O$_2$ - O unit, for LDA and gradient corrected calculations are given in Table~\ref{tab2}. {}From these results, it is clear that gradient corrections have only a minor effect on the relaxed equilibrium structure. As we have already noted for the perfect crystal case, modifications of atomic structure with respect to LDA results are more pronounced in the PW scheme. We have calculated the surface formation energy in the standard way, by subtracting from the slab total energy (18 atoms) three times the energy of a 6-atom perfect crystal unit cell and dividing by the total surface area. We find that the relaxed surface energy of SnO$_2$~(110) is 1.66~Jm$^{-2}$ in the LDA, 1.13~Jm$^{-2}$ in PW-GGC and 1.16~Jm$^{-2}$ for BP-GGC. The LDA result is close to the value of 1.50~Jm$^{-2}$ reported earlier~\cite{IanXX}. For TiO$_2$~(110), the values are 1.14, 0.82 and 0.84~Jm$^{-2}$ respectively. Our LDA result for TiO$_2$ (110) is close to the value of 1.06~Jm$^{-2}$ reported by Ramamoorthy {\em et al.}~\cite{ram94a}. Comparison of LDA and GGC results shows that gradient corrections have a substantial effect on the surface energies. For both GGC schemes, the surface energies are lowered by about 30\% with respect to the LDA values, the difference between PW and BP being very small. This decrease of surface energy by GGC is consistent with the general tendency of gradient corrections to remove the systematic overestimation of electronic binding energy in the LDA. The electronic DOS of the SnO$_2$~(110) surface using the LDA and the two GGC schemes are compared in Fig.~\ref{fig1}. In order to separate out effects of electronic structure, all the calculations are done at the equilibrium lattice parameters and the relaxed positions produced by the BP scheme. Overall, the differences between the three sets of results are small. However, there are significant differences at the top of the valence band and at the top of the O(2s) band. As we found in our previous work~\cite{IanYY} there are peaks at the top of both bands due to surface states, these states being concentrated on the bridging oxygens. The effect of GGC is to reduce the intensity of the peak at the top of O(2s) band. The effect on the intensity of the valence band peak is less systematic, since BP increases it but PW decreases it. The reason why these effects are interesting is that there appears to be no experimental evidence for the surface--state peak at the top of the valence band, so that the LDA predictions seems not to be consistent with experiment. The present results suggest the possibility that this inconsistency may be due to inaccurate treatament of exchange and correlation. \section{Conclusions} Our calculations show that gradient corrections increase the lattice parameters of TiO$_2$ and SnO$_2$ by $\sim$~4~\% for the Perdew-Wang scheme and $\sim$~3~\% for the Becke-Perdew scheme. These effects are similar to those reported previously for metals and semiconductors. For the surfaces we examined, gradient corrections have very little effect on the relaxed surface structure, but the surface energies are substantially reduced -- by $\sim$~30~\% in both the Perdew-Wang and the Becke-Perdew schemes. The effects of gradient corrections on the electronic DOS of SnO$_2$~(110) surface are very small, except at the top of the O(2s) and O(2p) bands. The changes we find at the top of the O(2p) band may be relevant to apparent inconsistences between calculated and experimental results for the surface DOS in this region. \section{Acknowledgments} The work of JG is supported by EPSRC grant GR/J34842. The major calculations were performed on the Intel iPSC/860 parallel computer at Daresbury Laboratory, and we are grateful for a generous allocation of time on the machine. Analysis of the results was performed using local hardware funded by EPSRC grant GR/J36266.
\section{Introduction} \label{sec:intr} In a recent paper \cite{KeIr95}, we demonstrated the power of finite--size scaling applied to Lee--Yang zeroes \cite{LY} in uncovering logarithmic corrections to scaling in the two--dimensional $XY$-- or $O(2)$ spin model. In this paper we apply the same techniques to the closely related \lq step model\rq{} \cite{GuJoTh72,GuJo73}, also known \cite{LeSh87,LeSh88} as the sgn $O(2)$ model. The question of criticality of this model has, until now, been unresolved despite several analyses based on high temperature series (see \cite{LeSh88} for a review) and on numerical simulation \cite{NyIr86,SVWi88}. The interest in the model arises from its possible membership of the $XY$--model universality class which exhibits the Kosterlitz--Thouless [KT] phase transition \cite{KT}. Like the $XY$--model, the step model has a configuration space which is globally and continuously symmetric. Unlike the $XY$--model, however, the interaction function is discontinuous and the Mermin--Wagner theorem \cite{MeWa66} does not apply. Nonetheless, it is expected that if a phase transition exists in the step model, it should not be to a phase with long range order \cite{GuJo73,SVWi88,GuNy78}. S\'anchez--Velasco and Wills \cite{SVWi88} presented evidence of critical behaviour starting at $\beta_c=1/T_c=0.91\pm0.04$. This was based on finite--size scaling [FSS] of the spin susceptibility. Since the associated critical index $\eta(T_c)$ was significantly greater than that measured for the $XY$--model, it was concluded that the step and $XY$ models are not in the same universality class. In this paper we present evidence that the step model is {\em not} critical at that temperature. However it {\em is} critical at lower temperatures with a critical index $\eta(T_c)$ compatible with the $XY$ value. The accuracy afforded by the Lee--Yang zeroes study is a crucial part of the analysis. \section{The step model and the $XY$--model} \label{sec:step} Consider the partition function \begin{equation} Z(\beta,h) = \sum_{\{{\vec{s}}_x\}}{ e^{- \beta H + h {\hat{n}}\cdot{\vec{M}} } } \, , \label{eqn:pf} \end{equation} where the Boltzmann factor is $\beta = 1/kT$, ${\hat{n}}$ is a unit vector defining the direction of the external magnetic field and $h$ is a scalar parameter representing its strength. The summation is over all configurations open to the system and ${\vec{s}}_x$ is a unit length two-component spin at each site $x$ in the cubic lattice $\Lambda\equiv L^d$ ($d=2$). The magnetisation for a given configuration is \begin{displaymath} {\vec{M}} = \sum_{x \in \Lambda} {\vec{s}}_x \, . \end{displaymath} In the case of the $XY$--model, the interaction hamiltonian is \begin{displaymath} H_{{XY}} = - \sum_{x \in \Lambda} \sum_{\mu = 1}^{d} {\vec{s}}_x\cdot {\vec{s}}_{x+\mu} \, , \end{displaymath} while for the step model it is \begin{displaymath} H_{\rm{step}} = - \sum_{x \in \Lambda} \sum_{\mu = 1}^{d} \hbox{sgn}( {\vec{s}}_x\cdot {\vec{s}}_{x+\mu}) \, . \end{displaymath} Thus the continuous (cosine) dependence of the interaction energy in the usual $XY$--model is replaced by a discrete step function dependence. The leading infinite volume critical behaviour of the 2D $XY$--model is characterized by exponential divergences (essential singularities) in the thermodynamic functions \cite{KT}. In terms of the reduced temperature $t\equiv 1-\beta/\beta_c \rightarrow 0^+$ the (leading) infinite volume scaling behaviour of the correlation length and the zero-field magnetic susceptibility is (respectively) \cite{KT} \begin{eqnarray} \xi_\infty(t) & \sim & e^{at^{-\nu}} \, , \label{eqn:ktxi} \\ \chi_\infty(t) & \sim & \xi_\infty^{2-\eta}\, , \label{eqn:ktchi} \end{eqnarray} where $\nu = 1/2$ and $\eta = 1/4$. The $XY$--model remains critical for all $ \beta > \beta_c$. For models obeying the Lee--Yang theorem \cite{LY}, the partition function zeroes in the magnetic field strength ($h$) plane (the Lee--Yang zeroes) are all on the imaginary axis. In the high temperature phase these zeroes remain away from the real axis, pinching it only as $t \rightarrow 0^+$ (in the thermodynamic limit). The zero closest to the real axis marks the edge of the distribution of zeroes and is known as the Yang--Lee edge, $z_{\rm{YL}}$. The theorem has been proved only for certain models, the $XY$--model included \cite{DuNe75}. In \cite{KeIr95} we used this fact to show the above leading critical behaviour for the $XY$--model and the corresponding behaviour of the Yang--Lee edge, $z_{\rm{YL}}$ are, in fact, modified by logarithmic corrections: \begin{eqnarray} \chi_\infty (t) & \sim & \xi_\infty^{2 - \eta} t^{r}\, , \label{eqn:chis} \\ z_{\rm{YL}}(t) & \sim & \xi_\infty^{\lambda} t^{p}\, , \label{eqn:edges} \end{eqnarray} where \begin{equation} \lambda= -\frac{1}{2}(d+2 -\eta) = -{{15}\over 8} \label{eqn:lambda} \end{equation} and the parameters $r$ and $p$ ($=-r/2)$ are logarithmic correction indices \cite{KeIr95}. The corresponding FSS behaviour for the susceptibility and the first zero $z_1$ ($=z_{YL}$) at $t=0$ is \cite{KeIr95} \begin{eqnarray} \chi_L (0) & \sim & L^{2 - \eta} (\ln{L})^{-\frac{r}{\nu}}\, , \label{eqn:chifss} \\ z_1(0) & \sim & L^{\lambda} (\ln{L})^{r} \label{eqn:edgefss} \, . \end{eqnarray} For the 2D $XY$--model the numerical value of $r$ was found to be small but non-zero ($-0.023\pm 0.010$) \cite{KeIr95}. The objects of the present analaysis were to establish \begin{enumerate} \item if the scaling behaviour of the very precisely determined Yang--Lee edge would unequivocably determine whether the step model had a critical phase \item if so, whether the phase transition is in the same universality class as the $XY$--model. \end{enumerate} \section{Method and results} \label{sec:meth} The methods used are those described in \cite{KeIr95}. A single cluster algorithm \cite{Wo89} is used to generate a large number of measurements (100K for each lattice size $L$ and temperature $1/\beta$) of the energy $H$ and the magnetisation ${\vec{M}}$ at zero external magnetic field. Such data were obtained for lattice sizes $L=32$, $48$, $64$, $128$ and $256$ covering the $\beta$ range $0.86$ to $1.40$ with varying degrees of spacing. Multi--histogram techniques \cite{FeSw88,KaKa91} were used to combine data at different values of $\beta$ and so obtain detailed $\beta$ dependence. In the neighborhood of possible critical points we used sufficiently fine spacing (typically 0.025) to ensure adequate overlap between histograms for a given size of lattice. The partition function for a complex magnetic field ($h=h_r+ih_i$, $h_r,h_i \in \mbox{\math R} $) can be written in terms of a real and an imaginary part \cite{KeIr95,KeLa93,KeLa94} as \begin{equation} Z(\beta,h_r + i h_i) = {\rm{Re}} Z(\beta,h_r + i h_i) + i {\rm{Im}} Z(\beta,h_r + i h_i) \, , \end{equation} where \begin{eqnarray} {\rm{Re}} Z(\beta,h_r + i h_i) & = & Z(\beta,h_r) \langle \cos{(h_i M)} \rangle_{\beta,h_r} \label{ReZ} \, , \\ {\rm{Im}} Z(\beta,h_r + i h_i) & = & Z(\beta,h_r) \langle \sin{(h_i M)} \rangle_{\beta,h_r} \label{ImZ} \, . \end{eqnarray} Here the subscripts indicate that the expectation values are taken at (inverse) temperature $\beta$ and in a (real) external field $h_r$. No specific proof of the Lee--Yang theorem exists for the step model. We can, however, use (\ref{ReZ}) and (\ref{ImZ}) to (numerically) determine the loci along which the real and imaginary parts of the partition function separately vanish. These formulae concern expectation values of real quantities in a real external field and at no stage is a simulation involving a complex action involved. The Lee--Yang zeroes are then the points in the complex $h$--plane where the loci intersect \cite{KeLa94}. We have determined these loci and thereby the Lee--Yang zeroes close to the real $h$--axis. We were able to determine the first 15 zeroes and found that they lie on the imaginary $h$--axis for all the lattices studied. Thus we have numerical evidence that the step model obeys the Lee--Yang theorem. The following analysis applies to the first zeroes only (the Yang--Lee edge) and we defer the study of higher zeroes to a later paper \cite{bbig}. The analysis began with a rough search for the leading critical behaviour predicted by (\ref{eqn:edgefss}). An independent test was also made using the (less accurate) susceptibility data and (\ref{eqn:chifss}). Both methods indicated critical behaviour setting in for $\beta$ \raisebox{-.75ex}{ {\small \shortstack{$>$ \\ $\sim$}} } $1.2$. In Fig.~\ref{fig:typscal} we show a typical log--log plot of $z_1$ the Yang--Lee edge versus $L$. This is at a typical candidate value of the critical temperature ($\beta=1.22$). The errors are considerably smaller than the symbols. For example, at $\beta=1.22$ we found $z_1=0.0024136(7)$ and $0.00017902(6)$ at $L=32$ and $128$ respectively. The slope of Fig.~\ref{fig:typscal} gives the effective leading index $\lambda_{\hbox{eff}}$ ignoring corrections. At $\beta=1.22$ this is $\lambda_{\hbox{eff}} = -1.8761(2)$ where the chi-squared per degree of freedom ($\chi^2/{\rm{dof}}$) for the linear fit shown is $0.85$. In Fig.~2(a) we display the result of such fits as a function of $\beta$. The effective exponent $\lambda_{\hbox{eff}}$ is just the slope of the log--log linear fit which should obtain if critical behaviour is present (ignoring logarithmic corrections). The corresponding $\chi^2/{\rm{dof}}$ for the linear fit is also shown. Acceptable values are only found for $\beta$ in excess of around 1.2. To quantify this statement we demand \begin{equation} \chi^2/{\rm{dof}} \leq 2.0 \label{eqn:chisq} \end{equation} which means $\beta \raisebox{-.75ex}{ {\small \shortstack{$>$ \\ $\sim$}} } 1.185$. We note that the corresponding values of $\lambda_{\hbox{eff}}$ ($ \raisebox{-.75ex}{ {\small \shortstack{$<$ \\ $\sim$}} } -1.872(2)$) include that ($-15/8 = -1.875$) corresponding to the KT prediction. Fig.~\ref{fig:lambda} is evidence for ({\em{i}}) the validity of FSS over a range of values $\beta \ge \beta_c \simeq 1.185$ and ({\em{ii}}) at $\beta_c$ the exponent $\lambda_{\hbox{eff}}$ very close to the expected KT value (-15/8). Observation ({\em{i}}) means that, as in the $XY$ case, the system remains critical for $\beta \ge \beta_c$ and ({\em{ii}}) is evidence that $\lambda$ is in fact -15/8 and the question of whether or not the step model belongs to the same universality class as the $XY$ model must now be answered by determination of the correction exponent $r$. We therefore {\em assume} the behaviour (\ref{eqn:edgefss}) with $\lambda = -15/8$ at $\beta=\beta_c$. The expected leading behaviour is removed and linear fits to \begin{equation} \ln \left( z_1 L^{15/8}\right) \quad\hbox{vs.}\quad\ln\ln L \end{equation} performed. The results are shown in Fig.~3. Since the value $\lambda=-15/8$ ($\eta=1/4$) is only expected at $\beta_c$ these results can be used to identify the possible values of critical temperature and to test for the presence of logarithmic corrections as in the $XY$--model \cite{KeIr95}. Applying the same criterion (\ref{eqn:chisq}) as for the leading behaviour, we search for a range of $\beta_c$ values giving an acceptable fit. We find \begin{equation} 1.195\leq\beta_c\leq 1.295\quad\hbox{and correspondingly,}\quad 0.009\geq r \geq -0.034\, . \end{equation} The range of acceptable $r$ values includes that found \cite{KeIr95} for the $XY$--model ($-0.023 \pm 0.010$) and that corresponding to no logarithmic corrections ($r=0$). As in our previous work \cite{KeIr95}, this range excludes the prediction $r = -1/16$ coming from an approximate renormalisation group treatment of the $XY$--model \cite{KT}. Thus we conclude that the present data are compatible with the step model being in the same universality class as the $XY$--model. We do not, however, exclude other possibilities. The susceptibility data are consistent with the above analysis. If one assumes the KT value of $\eta(\beta_c)=1/4$, one can construct a so-called Roomany--Wyld beta function approximant \cite{RW} from the finite--size data and use its zero to locate $\beta_c$ \cite{KeIr95}. These approximants, based on pairs of lattice size $L, L'$, converge very rapidly \cite{RW}. We estimate $\beta_c=1.22 \pm 0.02$. This last analysis of course neglects possible logarithmic corrections to scaling. We have also studied the specific heat. As for the $XY$--model, the step model data show a broad peak with no obvious relationship to the position of the leading critical point. The finite--size dependence is not dramatic and is likely to be of little value in further elucidating the criticial behaviour. A related question is to what extent one can make use of the Fisher zeroes \cite{KeLa93,Fi72}, i.e. zeroes of the partition function in the complex $\beta$ plane at zero external magnetic field $h$. For both this and the $XY$--model, these are much harder to locate than the Lee--Yang zeroes and are consequently less accurately determined. \section{Conclusions} \label{sec:conc} The use of finite--size scaling applied to Lee--Yang zeroes has allowed us to present detailed evidence of critical behaviour in the two-dimensional step (sgn $O(2)$ spin) model. The data are consistent with this model being in the same universality class as the $XY$-- ( $O(2)$ spin) model. That is, it undergoes a Kosterlitz--Thouless type transition with susceptibility index $\eta({\beta_c})=1/4$ and we determine that $\beta_c$ lies in the range $1.195 \leq \beta_c \leq 1.295$. With the available statistics, we found the logarithmic correction exponent to lie in the range $ -0.034 \leq r \leq 0.009$. This should be compared with our measurement for the $XY$ correction exponent \cite{KeIr95}, $-0.033 \leq r \leq -0.013$ with which it is compatable. The step model results are however also compatable with no logarithmic corrections ($r=0$ corresponds to $\beta_c = 1.21$ and $\chi^2/{\rm{dof}}= 0.92$). The Mermin--Wagner theorem \cite{MeWa66} does not apply directly to the step model because of the discontinuous nature of the interaction hamiltonian. However, it has long been believed \cite{GuJo73,SVWi88} that if a phase transition does exist, it will not involve a phase with long range order. The evidence presented here supports this view. This raises a question as to the nature of the mechanism driving the phase transition in the step model. The KT phase transition of the $XY$--model is understood to be driven by the binding/unbinding of topological solutions (vortices). However, the energetics of vortex formation are very different in the step model \cite{GuNy78,LeSh88,SVWi88}. Since vortices with effectively zero excitation energy can be created at all non-zero temperatures, the usual KT argument does not naturally lead one to expect such a phase transition in the step model. If this is indeed the case, some other driving mechanism must be responsible for any phase transition. It would then be remarkable if --- as the evidence presented here indicates -- both models belong to the one universality class.
\section{Introduction} Black holes seem to be a never-ending source of surprises. While much has been learned about their behavior, much remains to be understood -- even at the classical level. In this paper we study the classical supersymmetric solutions in a general theory with N = 2 supersymmetry. Previous work on this subject can be found in \cite{GH,T}. The solutions appear to have a richer structure than the more thoroughly studied N = 4 case. In section 2 we recall some aspects of N = 2 supergravity. In section 3 the magnetic solutions are described in terms of trajectories in the special geometry of the N = 2 moduli space which terminate at a supersymmetric fixed point at the horizon. In section 4 we find that the equations can be integrated for a restricted but large class of cases. An intriguing relation between the K\"{a}hler potential on the moduli space and the metric conformal factor emerges. Some simple examples are worked out in detail in section 5. We do not attain a complete characterization of the classical geometry of N = 2 black holes in this paper, but we hope that our results prove useful for future efforts in this direction. \section{Special Geometry and N = 2 Supersymmetry} We study N = 2 supergravity coupled to $n$ \ N = 2 vector multiplets in the framework of special geometry \cite{CREM}--\cite{CAFP}. In this section some formulae that will be needed in the following are recalled. Further details can be found in \cite{WLP} whose notation we adopt. The supergravity theory is defined in terms of a projective covariantly holomorphic section $(X^\Lambda(\phi^i), -{i\over 2}F_\Lambda(\phi^i))$, $\Lambda = 0, 1, ..., n, ~~i=1,...,n$, of an ~$Sp(2n+2)$ vector bundle over the moduli space parametrized by $\phi^i$. (We note that alternate conventions are often employed in which the definition of $F_\Lambda$ differs by a factor of $2i$.) In some cases the theory can be described in terms of a covariantly holomorphic function $F(X)$ of degree two: \begin{equation}\label{2} F_\Lambda(\phi^i) = F_\Lambda(X(\phi^i)) = {\partial\over \partial X^\Lambda} F(X) \ . \end{equation} Given a prepotential $F(X)$, or a covariantly holomorphic section $(X^\Lambda,\, -{i \over 2} F_\Lambda)$, one can construct the entire scalar and vector parts of the action. It is convenient to introduce the inhomogeneous coordinates \begin{equation}\label{10} Z^\Lambda = {X^\Lambda(\phi_i)\over X^0(\phi_i)} \ , ~~~~~~ Z^0 = 1 \ . \end{equation} We assume $Z^\Lambda(\phi_i)$ to be invertible, so that, in special coordinates, ${\partial Z^\Lambda\over \partial \phi^i} =\delta^\Lambda_i$. ~In this case the complex scalars $Z^i = \phi^i$ {}~($ i = 1,...,n$) represent the lowest component of the $n$ vector multiplets of N = 2 supersymmetry. The K\"{a}hler potential determining the metric of these fields is \begin{eqnarray}\label{22} &&K(Z,\bar Z)=2\ln |X^0| = -\ln \left (N_{\Lambda\Sigma}(Z,\bar Z)\, Z^\Lambda\bar Z^\Sigma\right )\nonumber \\ &&= - \ln {1\over 2} [f(Z) + \bar f(\bar Z) + {1\over 2} (Z^i - \bar Z^i) (\bar f_i - f_i)] \ , \end{eqnarray} where $N_{\Lambda\Sigma} = {1\over 4}(F_{\Lambda\Sigma} + \bar F_{\Lambda\Sigma})$ and $f(Z) = (X^0)^{-2} F(X)$. In the conformal gauge \cite{WLP} \begin{equation}\label{conf} N_{\Lambda\Sigma} X^\Lambda \bar X^{\Sigma}=1 \ . \end{equation} The graviphoton field strength, as well as the field strengths of the $n$ Abelian vector multiplets, are constructed out of $n+1$ field strengths $\hat F^\Lambda_{\mu\nu} = \partial_\mu W^\Lambda_\nu - \partial_\nu W^\Lambda_\mu$. The graviphoton field strength is \begin{equation}\label{23} T_{\mu\nu}^{+} = {4 N_{\Lambda\Sigma} X^\Lambda\over N_{IJ} X^I X^J} \ \hat F_{\mu\nu}^{+\Sigma} \ , \end{equation} where the superscript + (-) denotes the (anti)-self-dual part. This defines the central charge of the theory, since it enters into gravitino transformation rules. The vector field strengths which enter the gaugino supersymmetry transformations are \begin{equation}\label{24} {\cal F}_{\mu\nu}^{+\Lambda} = {\hat F}_{\mu\nu}^{+\Lambda} -{1\over 4} X^\Lambda\, T_{\mu\nu }^{+} \ . \end{equation} The anti-self-dual vector field strengths ${\hat F}^{-\Lambda}$ are part of the symplectic vector \begin{equation}\label{11} \pmatrix{ {\hat F}^{-\Lambda} \cr -2i \bar { \cal N}_{\Lambda\Sigma}\,{\hat F}^{-\Sigma} \equiv i G^-_{\Lambda} \cr } , \end{equation} where again the expression for the matrix ${\cal N}_{\Lambda\Sigma}$ is derived from the prepotential $F(X)$ \cite{WLP}--\cite{CAFP}. The vector part of the action is then proportional to \begin{equation}\label{12} \mbox{ Re}\ {\hat F}^{-\Lambda} G^-_\Lambda \ , \end{equation} and the graviphoton field strength can be written in the manifestly symplectic form \begin{equation}\label{tst} T_{\mu\nu}^- = 2X^\Lambda G^-_{\Lambda\mu\nu}+ F_{\Lambda} {\hat F}_{\mu\nu}^{-\Lambda} \ . \end{equation} The Lagrangian for the scalar components of the vector multiplets is defined by the K\"{a}hler potential as \begin{equation}\label{13} g_{i\bar j}\,\partial_\mu\phi^i\, \partial_\nu\bar\phi^{\bar j}\, g^{\mu\nu} \ , \end{equation} where $g^{\mu\nu}$ is the space-time metric, and \begin{equation}\label{14} g_{i\bar j}= \partial_i \partial_{\bar j} K(\phi, \bar\phi) \ . \end{equation} The gravitino supersymmetry transformation law, to leading order in fermi fields, is \begin{equation}\label{grt}\delta \psi^\alpha _\mu=2\nabla_\mu\epsilon^\alpha -{1 \over 16}\gamma^{\nu\lambda}T^-_{\nu\lambda}\gamma_\mu \epsilon^{\alpha \beta} \epsilon_\beta+iA_\mu\epsilon^\alpha \ , \end{equation} where $ \alpha, \beta=1,2$ are $SU(2)$ indices and $A_\mu= {i \over 2} N_{\Lambda\Sigma} [\bar X^\Lambda \partial_\mu X^\Sigma - (\partial_\mu \bar X^\Lambda) X^\Sigma]$. The gaugino transformation law is \begin{equation}\label{ggt}\delta \Omega^\Lambda_\alpha =2\gamma^\mu\nabla_\mu X^\Lambda \epsilon_\alpha +{1 \over 2}\gamma^{\nu\lambda} {\cal F}^{+\Lambda}_{\nu\lambda} \epsilon_{\alpha \beta} \epsilon^\beta+2i\gamma^\mu A_\mu\epsilon_\alpha \ . \end{equation} BPS states of the N = 2 theory have a mass equal to the central charge $z$. It follows from the supersymmetry transformation rules that this is simply the graviphoton charge \cite{CAFP} \begin{equation}\label{15} M = |z| = |q^{(e)}_\Lambda X^\Lambda - {i \over 2} q_{(m)}^\Lambda F_\Lambda|= e^{K/2} |q^{(e)}_0 + q^{(e)}_i Z^i + {i \over 2} (q_{(m)}^0 Z^i - q_{(m)}^i)f_i - iq^0_{(m)} f| \ , \end{equation} where $q^{(e)}$ and $q_{(m)}$ are electric and magnetic charges associated to $i\cal G$ and $\hat F$ and comprise a symplectic vector. Duality transformations of the N = 2 theory correspond to different choices of the symplectic representative $(X^\Lambda,-{i \over 2}F_\Lambda)$ of the symplectic geometry. \section{ Magnetic N = 2 BPS Black Holes} In this section we discuss the general form of the supersymmetric magnetic black hole solutions and their interpretation as interpolating solitons. It has been shown by Tod \cite{T} in general that, for N = 2 theories, a static metric admitting supersymmetries can be put in the form \begin{equation}\label{17} ds^2 = -e^{2U} dt^2 + e^{-2U} d\vec x^2 \ . \end{equation} For spherically symmetric black hole solutions $U$ will be a function only of the radial coordinate $r$. By solving the Bianchi identities $d \hat F^\Lambda = 0$ we then find for the radial component of the magnetic\footnote{ It follows from equation (7) that for a generic prepotential $F$ the electric charge will be nonzero despite the fact that $\hat F$ is a magnetic field strength. In general this electric charge will not respect the charge quantization condition. This can be avoided by restricting the prepotential as in the examples in section 5. } field strengths $ \hat F^\Lambda_r \equiv 2{\epsilon_r}^{\theta \phi} \hat F^\Lambda_{\theta \phi}$ : \begin{equation}\label{26} \hat F^\Lambda_r = {q^\Lambda\over r^2} \, e^{U(r)} \ . \end{equation} Inserting (\ref{17}) and (\ref{26}) into the gravitino transformation law, and demanding that the variation vanish for some choice of $\epsilon$, we derive the following first order differential equation: \begin{equation}\label{27} 4 U'= -\sqrt{(\bar ZNq) (ZNq) (\bar Z N Z)\over (ZNZ) (\bar Z N \bar Z)}~e^{U}\ , \end{equation} where $U' \equiv {\partial U\over \partial \rho}$, $\rho\equiv 1/r$, and we employ the notation $(ZNq)\equiv Z^\Lambda N_{\Lambda\Sigma}\,q^\Sigma$. Equation (\ref{27}) may be viewed as determining $U$ as a function of the moduli fields $Z^\Lambda$. A vanishing \ gaugino transformation further requires that the moduli fields obey \begin{equation}\label{28} (Z^\Lambda)' = -{e^{U}\over 4}\, \sqrt{(ZNZ) (\bar Z Nq) (\bar Z NZ) \over (\bar Z N \bar Z) (ZNq)}~ \Bigl( Z^\Lambda q^0 - q^\Lambda\Bigr) \ . \end{equation} Differentiating again with respect to $\rho$ and substituting (\ref{27}) leads to the second order differential equation \begin{equation}\label{28g} (Z^\Lambda)''-\left( { (ZNq) \over (ZNZ) }+q^0 \right) {((Z^\Lambda )')^2 \over Z^\Lambda q^0 - q^\Lambda} +{1 \over 2} \left(\ln {(ZNZ) (\bar Z Nq) (\bar Z NZ) \over (\bar Z N \bar Z) (ZNq)} \right)' (Z^\Lambda )'=0\ . \end{equation} This equation is independent of $U$ and can be viewed as a generalized geodesic equation which describes how $Z$ evolves as one moves into the core of the black hole. Initial conditions for $Z$ are specified at infinity ($\rho=0$) corresponding to the asymptotic values of the field. The first derivative of $Z$ is then fixed in terms of the charge of the black hole by the supersymmetry constraint (\ref{28}). $Z$ will then evolve until it runs into a fixed point. It is evident from (\ref{28}) and (\ref{28g}) that these fixed points are at \begin{equation}\label{zf} Z^\Lambda_{fixed}={q^\Lambda \over q^0} \ , \end{equation} where $(Z^\Lambda)'=0$. Each fixed point is typically surrounded by a finite basin of attraction. The limiting form of the metric at such a fixed point is found by integrating (\ref{27}), \begin{equation}\label{uc} e^{-U}={c \over 4}\, \rho \ , \end{equation} where the constant $c$ is given by \begin{equation}\label{cv} c ={\sqrt{ q^\Lambda N_{\Lambda\Sigma}\,q^\Sigma }} =q^0\, e^{-K(Z^\Lambda_{fixed} )/2}\ . \end{equation} This corresponds to the maximally symmetric charged Robinson-Bertotti universe. Thus, as in \cite{GT}, the extremal black holes may be viewed as solitons which interpolate between maximally symmetric vacua at infinity and the horizon. The locations of the fixed points (\ref{zf}) depend on the charges but not on the asymptotic values of the moduli fields. Thus if the the asymptotic values of those fields are adiabatically changed, the geometry of the black hole near the horizon remains fixed. Symplectic invariance implies a similar structure for many of the dyonic and electrically charged extremal black holes. \section{Space-Time Geometry from K\"{a}hler Geometry} In this section we consider a remarkably simple special class of solutions which exist for a generic prepotential. We will work in a symplectic basis in which $q^0=0$ in order to simplify the equations. In such a basis the fixed points $Z^\Lambda_{fixed}$ move to infinite coordinate values. Thus solutions for which the moduli field is constant at the fixed point (which corresponds to the Reissner-Nordstr\"om solution) cannot be described in this basis\footnote{However, by using the manifestly symplectic constraint \cite{CAFP} instead of the superconformal one \cite{WLP}, one can describe the Reissner-Nordstr\"om configuration in terms of the K\"{a}hler potential, see example 5.4.2}. In the $q^0=0$ basis it is straightforward to check that (\ref{27}) and (\ref{28}) are solved by \begin{equation}\label{31} e^{2U(\rho)}= e^{K(Z,\bar Z)- K_\infty} \ , \end{equation} and \begin{equation}\label{harmreal} Z^i = Z^i_{\infty} + {q^i \over 4} \rho \, e^{-K_\infty/2} \ , \end{equation} provided that the asymptotic value of $Z^i$ is restricted to obey \begin{equation}\label{hrm} Z^i_{\infty} = \bar Z^i_{\infty} \ . \end{equation} Alternatively we may have \begin{equation}\label{harmimagin} Z^i = Z^i_{\infty} + i {q^i \over 4} \rho \, e^{-K_\infty/2} \ , \end{equation} provided that the asymptotic value of $Z^i$ is restricted to obey \begin{equation}\label{hrm2} Z^i_{\infty} = - \bar Z^i_{\infty} \ . \end{equation} The restrictions imply that these solutions exist only at special points in the moduli space where all $Z^i$ are either real or imaginary. More general solutions may be obtained from these by symplectic transformations. Thus the space-time metric is \begin{equation}\label{34} ds^2 = e^{K(Z^i,\bar Z^i=Z^i) - K_\infty}\, dt^2 - e^{-K(Z^i,\bar Z^i=Z^i) + K_\infty}\, d\vec x^2 \ , \end{equation} where each $Z^i$ solves a three-dimensional harmonic equation and is given in eq. (\ref{harmreal}) or (\ref{harmimagin}). Hence the logarithm of the spatial conformal factor is identified with the moduli space K\"{a}hler potential. \section{Examples of\, ${\bf N \geq 2}$\, BPS states} \subsection{ Calabi-Yau magnetic black holes} The prepotential is \begin{equation}\label{44} F = i d_{ABC}\, {X^A X^B X^C \over X^0} \ . \end{equation} We consider pure imaginary $Z^A$ and real $d_{ABC}$ (corresponding to the classical Calabi-Yau moduli space) \begin{equation} \qquad e^{- K(Z,\bar Z)} = - 2 d_{ABC} \, {\rm Im} Z^A \, {\rm Im} Z^B \, {\rm Im} Z^C, \end{equation} where ${\rm Im} Z^A = {1\over 2i} (Z^A - \bar Z^A)={\rm Im} (Z^A)_\infty + ~ {q^A_{(m)}\over r}\, e^{-K_\infty/2}$. \begin{equation}\label{45} ds^2 = \left( { \, d_{ABC} \, {\rm Im} Z^A \, {\rm Im} Z^B \, {\rm Im} Z^C \over \, [d_{ABC} \, {\rm Im} Z^A \, {\rm Im} Z^B \, {\rm Im} Z^C]_\infty }\right )^{-1}\, dt^2 - \left( { \, d_{ABC} \, {\rm Im} Z^A \, {\rm Im} Z^B \, {\rm Im} Z^C \over \, [d_{ABC} \, {\rm Im} Z^A \, {\rm Im} Z^B \, {\rm Im} Z^C]_\infty }\right )\, d\vec x^2 \ . \end{equation} \subsection{Massive and massless ${\bf SU(1,n) \over SU(n)} $ supersymmetric white holes} The prepotential is \begin{equation}\label{36} F(X^0,X^1) = (X^0)^2- (X^i)^2 \ , \qquad e^{- K(Z,\bar Z)} = 1-|Z^i|^2 \ . \end{equation} Here the $Z^i$ are real: \begin{equation}\label{38} Z^i = Z_\infty^i + {q^i \over r}\, e^{-K_\infty/2} \ , \end{equation} \begin{equation}\label{39} ds^2 = \left({1-|Z|^2\over 1-|Z_\infty|^2}\right)^{-1}\, dt^2 - \left({1-|Z|^2\over 1-|Z_\infty|^2}\right)\, d\vec x^2 \ . \end{equation} In particular in the simplest case of $i=1$ we get \begin{equation}\label{40} g_{tt} = g_{rr}^{-1} = \left(1- {2 Z_\infty\, q\over r}\,e^{K_\infty/2} - {q^2\over r^2}\right)^{-1} . \end{equation} To satisfy the supersymmetry bound we require \begin{equation}\label{41} M =- {Z_\infty \, q\over\sqrt{1-|Z_\infty|^2}} \geq 0 \ , \end{equation} and the geometry is \begin{equation}\label{42} g_{tt} = g_{ii}^{-1} = \left(1+{2 M\over r} - {q^2\over r^2}\right)^{-1}. \end{equation} These configurations are non-trivial in the limit when the ADM mass tends to zero. It will be interesting to understand if or when such states can arise in a physical theory. The limit $M \to 0$ can be achieved via $Z_\infty \to 0$. In this limit the scalar field becomes inversely proportional to the radius $r$, \begin{equation}\label{43} Z(r) = {q\over r} \ . \end{equation} For N = 4 BPS states the analogous massless states have been studied recently \cite{Klaus2}--\cite{CYHETS}. The configuration exhibits a repulsive (i.e. antigravitating) singularity and was referred to as a supersymmetric white hole \cite{KL}. The difference between the metric (\ref{42}) for N= 2 and the corresponding metric obtained in \cite{Klaus2}--\cite{KL}\, for N = 4 is that $g^{N = 4}_{tt} = (g^{N = 2}_{tt})^{1/2}$. This does not change the repulsive nature of the singularity. \subsection{$\bf { SO (2,1)\over SO(2) } \times {SO(2,n) \over SO(2) \times SO(n)}$\, BPS states} Here again we consider pure imaginary $Z$. \begin{equation}\label{48} F = -i\, {X^s \over X^0}\, \Bigl((X^1)^2 - \sum^n_{a = 2}(X^a)^2 \Bigr) \ , \qquad e^{- K(Z,\bar Z)} = 2 {\rm Im} Z^s\, \Bigl(( {\rm Im} Z^1)^2 - \sum^n_{a= 2}( {\rm Im} Z^a)^2 \Bigr)\ . \end{equation} The metric of the BPS configuration defined by the K\"{a}hler potential is given by \begin{equation}\label{49} g_{tt} = g^{-1}_{ i i} = \left[{ {\rm Im} Z^s \Bigl(({\rm Im} Z^1)^2 - ({\rm Im} Z^a)^2\Bigr)\over \left[ {\rm Im} Z_s \Bigl(({\rm Im} Z^1)^2 - ({\rm Im} Z^a)^2\Bigr)\right]_{\infty} }\right]^{-1}. \end{equation} The configurations presented in this example may contain both types of supersymmetric states, those with attractive singularities and those with the repulsive ones, depending on the choice of the parameters describing the harmonic functions \begin{equation}\label{46a} Z^i = Z^i_\infty + ~ i {q^i_{(m)}\over r}\,e^{-K_\infty/2}\ ,~~~~~ i = \{ s, 1, a=2, \dots , n \} . \end{equation} Note that this example actually provides one of the particular choices of Calabi-Yau magnetic black holes. \subsection{$\mbox{N=4, 2}$ pure supergravity black holes from K\"{a}hler geometry perspective} Here we will analyse some previously known\, supersymmetric black hole solutions in the framework of the manifestly symplectic formalism \cite{CAFP}. The K\"{a}hler potential is different from the conformal gauge (\ref{conf}) (here $X_\Lambda, \, F_\Lambda $ are holomorphic sections) and is given by \begin{equation}\label{a7} e^{- K(X, \bar X) } = \bar X^\Lambda N_{\Lambda \Sigma} X^\Sigma \ . \end{equation} \subsubsection{$\mbox {SL(2,Z)}$ axion-dilaton dyons} It is instructive to analyse the known $SL(2,Z)$-invariant axion-dilaton dyonic black hole solution \cite{KO} of N = 4 supergravity as a particular solution of N = 2 supergravity coupled to an N = 2 vector multiplet. These solutions have complex moduli. The prepotential for this theory is $F(X) = 2 X^0 X^1$. The holomorphic section includes \begin{equation}\label{a4} \left (\matrix{ X^\Lambda\cr -{i\over 2} F_\Lambda \cr }\right ) \Longrightarrow \left (\matrix{ X^0\cr -iX^1\cr }\right ), ~~ \left (\matrix{ X^1\cr -iX^0 \cr }\right ). \end{equation} For the axion-dilaton black hole $X^0$ and $X^1$ can be identified with two complex harmonic functions ${\cal H}_1$, ${\cal H}_2$ as follows: \begin{equation}\label{a6} X^1 = {\cal H}_1(\vec x) \ , ~~~~~ X^0 = i {\cal H}_2(\vec x) \ . \end{equation} The K\"{a}hler potential is \begin{equation}\label{a7a} e^{- K(X, \bar X) } = i \,(\bar {\cal H}_1 {\cal H}_2 -\bar {\cal H}_2 {\cal H}_1) = g_{tt}^{-1}(\vec x) \ , \end{equation} in agreement with the metric of the axion-dilaton black hole found in \cite{KO}. \subsubsection{Reissner-Nordstr\"om solution} The prepotential for pure $N=2$ supergravity is $F(X) = ( X^0 )^2$. The K\"{a}hler potential of this theory in the manifestly symplectic formalism \cite{CAFP} is given by \begin{equation}\label{a7aa} e^{- K(X, \bar X) } = X^0 \bar X^0 = V^{-1} (\vec x ) \bar V^{-1} (\vec x ) =e^{-2U (\vec x)}\ , \end{equation} where $X^0=V^{-1} (\vec x )$ is a real (imaginary) harmonic function for electric (magnetic) Reissner-Nordstr\"om extremal supersymmetric black hole \cite{GH,T}. \vskip .5 cm In conclusion, we have found a simple relation between the special geometry describing the couplings of scalars and vectors in extended locally supersymmetric theories and the space-time geometry of the black-hole-type solutions in these theories. It is likely that more general solutions with complex moduli will be found in this framework. \section*{Acknowledgements} We are grateful to the Aspen Center for Physics where this work was initiated. S.F. was supported by DOE grants DE-AC0381-ER50050 and DOE-AT03-88ER40384,Task E, and by EEC Science Program SC1*CI92-0789.\, R.K. was supported by NSF grant PHY-8612280.\, A.S. was supported by DOE grant DOE-91ER40618.
\section{Introduction} \hspace{0.8cm} Particles in the expanding Robertson-Walker Universe are decelerated in the comoving frame. Ignoring backreaction, they obey the geodesic equation and lose their physical momentum $P_{phys}$ as \begin{equation} P_{phys} = \frac{P_{conf}}{a} \rightarrow 0, \end{equation} where $P_{conf}$ is a conserved conformal momentum and $a$ is a scale factor growing in time. In general particles in deceleration can emanate radiation, or some massless particles. We call this process geometric bremsstrahlung\footnote{ This process is also called by DeWitt and Brehme electro-gravitic bremsstrahlung in ref \cite{DB}.} due to the cosmic expansion. Many analyses on phenomena in the early Universe have been performed so far using results of high energy particle physics and proposed a lot of interesting features of the Universe\cite{KT}. However they are based on calculations of transition matrices in the flat spacetime, emphasizing the fact that rates of interactions are much larger than expansion rate given by Hubble parameter, and no careful attention seems to be paid to the geometric bremsstrahlung process, which actually gives no contribution in the flat spacetime. We shall argue in this paper that geometric bremsstrahlung may play impotant roles in the early Universe. In the section 2, emission of electromagnetic wave from a classical charged particle in the expanding Universe is discussed, taking account of backreaction. Our argument in the classical level suggests that the emission rate may not be simply ignored and the damping time may nearly equal to expansion time. Thus in the section 3, we treat photon emission from charged a particle quantum mechanically. High momentum limit of the transition probability can be obtained analytically. We stress that massless limit should be treated carefully and is nontrivial. In this paper, we adopt the natural units, the light velocity $c=1$ and the Planck constant $\hbar =1$. Signature of metric is taken as $(+,-,-,-)$. \section{Classical Geometric Bremsstrahlung in the Early Universe} \hspace{0.8cm}The radiation reaction has been neglected in the study of the early Universe. However particles are deaccelerated due to the cosmic expansion and thus they will emit the radiation. If the damping time due to the radiation reaction is comparable to the expansion time, the effect of the radiation reaction may not be simply neglected. We shall study in detail this phenomenon in the case of classical charged particles. The study of the radiation reaction for a charged particle has a long history. The first relativistic calculation was performed by Dirac\cite{Dirac}. His calculation has been generalized by DeWitt-Brehme\cite{DB} for the motion in gravitational field. They have shown that bremsstrahung induced by the spacetime curvature which we call geometric bremsstrahlung occured in addition to the usual radiation damping. The effect is nonlocal in general which is caused by the so-called tail term in the Green function. It was Hobbs\cite{Hobbs} who corrected the result of De Witt-Brehme and pointed out that the tail term vanishes identically in the case of the conformally flat spacetimes. His equation of motion for a particle with 4 velocity $ u^\mu $, mass $m$, charge $e$ without external electromagnetic field may be written in the following form in conformally flat spacetime. \begin{eqnarray} m \frac{D u^\mu}{D\tau} = \frac{2 e^2}{3} \left( \frac{D^2 u^\mu}{D \tau^2} + u^\mu \left( \frac{D u}{D\tau} \right)^2 \right) +\frac{2e^2}{3} ( \Omega_{,\alpha\beta} - \Omega_{,\alpha} \Omega_{,\beta} ) \left( g^{\mu\alpha} u^\beta - u^\mu u^\alpha u^\beta \right) \label{20} \end{eqnarray} where $D/D\tau $ is the absolute derivative along the worldline of the particle with $ \tau$ the proper time and the $ \exp (2 \Omega ) $ is the conformal factor, $ g_{\mu\nu} = e^{2\Omega} \eta_{\mu\nu} $ with $ \eta_{\mu\nu} $ the flat Minkowskii metric. Here we are interested in the radiation reaction induced by the cosmic expansion in the early Universe and thus we will restrict ourselves to the case where the conformal factor depends only on the time variable. We shall take a different approach from Hobbs and evaluate explicitly the damping time scale due to the geometric bremsstrahlung in the early Universe. \bigskip We shall take the standard form for the action of a charged particle with mass $m$ and charge $e$ in the gravitational field, $$ S = - m \int \sqrt{g_{\mu\nu} {\dot{x}}^\mu {\dot{x}}^\nu } d\tau - e \int A_\mu \dot{x}^\mu d\tau - {1\over 4} \int d^4 x \sqrt{-g} g^{\alpha\beta} g^{\mu\nu} F_{\alpha\mu} F_{\beta\nu} . $$ The dot denotes the derivative with respect to the proper time $\tau$. The equation of motion derived from the above action may be written as follows. \begin{eqnarray} m {{D u^\mu}\over {D\tau}} = m \left( {{du^\mu}\over {d\tau}} + \Gamma^\mu_{\alpha\beta}u^\alpha u^\beta \right) = e F^{\mu\nu} u_\nu.\nonumber \end{eqnarray} Since we are interested in the early Universe, we may neglect the spatial curvature and thus take the spatially flat Robertson-Walker model as our background geometry, $$ ds^2 = a^2 (\eta) ( d\eta^2 - d{\vec{x}}^{\ 2} ) =dt^2 -a^2 d{\vec{x}}^{\ 2} . $$ Since the metric is conformally flat, it is convenient to work in the conformally related flat spacetime. Defining the conformally related proper time $ d\tau_f = a^{-1} d\tau $, we shall define the conformally related 4 velocity $$ {\tilde u}^\mu = {{dx^\mu}\over {d\tau_f}} . $$ Then the equation of motion may be written as follows. $$ m {{D u^\mu}\over {D\tau}} = m a^{-2} \left( {{d{\tilde u}^\mu}\over {d\tau_f}} + {{a'}\over a} {{d\eta}\over {d\tau_f}} {\tilde u}^\mu - {{a'}\over a} \delta^\mu_0 \right) = e a^{-3} \eta^{\mu\alpha} F_{\alpha\beta} {\tilde u}^\beta . $$ where the prime denotes the derivative with respect to the conformal time $\eta $. By using the self field of the particle in the right hand side of the above equation, we shall obtain the radiation reaction force. Before calculating the reaction force explicitely, let us compare the timescale due to the radiation damping with that due to the cosmic expansion to see the importance of the radiation reaction in the early Universe. The damping time may be roughly evaluated as follows. $$ {1\over {t_{r}}} \sim \left| {1\over E_{conf}} {{dE_{conf}}\over {ad\eta}} \right| = {1\over p_{conf}} \, {2\over 3 a} e^2 \left({{d{\tilde{u}} }\over {d\eta}}\right)^2 =\frac{2e^2}{3p_{conf} a} \left(\frac{p_{conf} }{m a^2} \frac{da}{d\eta} \right)^2 = {2\over 3} e^2 \, {{p_{phys} H^2}\over {m^2}} $$ where $$ H= \frac{1}{a}\frac{da}{dt} $$ is the Hubble parameter, $p_{conf}$ is the conformal momentum and we have used the fact that the physical momentum $ p_{phys}= m{\tilde{u}}=p_{conf}/a $ decays as $ a^{-1}$ if the radiation reaction is neglected. Thus the ratio between the Hubble time $ t_{exp} = H^{-1} $ and the damping time is $$ {{t_{exp}}\over {t_{r}}} \sim {{2e^2}\over {3m^2}} p_{phys} H . $$ This ratio is much larger than the unity for a relativistic particle at sufficiently early times in the Universe. Thus the radiation reaction may not be simply ignored and might play an important role in the early Universe. \bigskip For the calculation of the reaction force, we shall need the field equation derived from the above action, $$ \eta^{\mu\nu} \eta^{\alpha\beta} F_{\alpha\nu,\beta} = e \int d \tau_f \delta^4 ( x - x( \tau_f)) {\tilde u}^\mu . $$ Taking the following non-covariant gauge $$ \eta^{\mu\nu} A_{\mu,\nu} = 0, $$ we arrive at the field equation which has the same form with that in the flat spacetime, $$ \eta^{\alpha\beta} A^\mu_{,\alpha\beta} = e \int d \tau_f {\tilde u}^\mu \delta^4 ( x - x(\tau_f)) . $$ Then the calculation by Dirac \cite{Dirac} applies here and we obtain the standard expression for the reaction force in the flat spacetime, $$ F^{\mu}_{react} =e\ \eta^{\mu\alpha} F_{\alpha\beta} {\tilde u}^\beta = {2\over 3} e^2 \left[ {{d^2 {\tilde u}^\mu}\over {d\tau_f}^2} + \left( {{d{\tilde u}}\over {d\tau_f}}\right)^2 {\tilde u}^\mu \right] . $$ It can be shown by a direct calculation that our expression of the equation of motion with radiation reaction in the conformally related flat spacetime coincides with eqn(\ref{20}) when transformed back to the original physical frame. \bigskip In order to see the effect of the radiation reaction explicitly, we shall forcus our attention to an 1-dimensional motion. Then the above equation is simplified as $$ {d\over {d\tau_f}} ( a m {\tilde u} ) = {2\over 3} e^2 \left[ {{d^2{\tilde u}}\over {d\tau_f}^2} - {{\tilde u}\over { 1 + {\tilde u}^2}} \left({{d{\tilde u}}\over {d\tau_f}}\right)^2 \right] . $$ Without the radiation reaction, the conformal momentum $ p_{conf} = am{\tilde{u}} $ is conserved as expected. Now we shall rewrite the above equation using the conformal momentum $p_{conf}$ and the background time $ dt = a d\eta $, \begin{eqnarray} {{d^2 p_{conf}}\over {dt^2}} = \left( H + {{3m}\over {2e^2 \sqrt{ 1 + (p_{phys}/m)^2}}} \right) {{d p_{conf}}\over {dt}} + {{dH}\over {dt}} p_{conf} . \label{21} \end{eqnarray} Notice that there will be no classical geomotric bremsstrahlung in the case of de Sitter expansion, namely $H= const$. We shall be interested in the relativistic case in the early universe, namely $$ p_{phys} \gg {{3m^2}\over {2e^2 H}}, \; \; \; m . $$ Then the second term in the coefficient of $ dp_{conf}/dt $ in eqn(\ref{21}) is negligible. Thus when the particle is relativistic, its evolution is governed by the reaction force only and the Hubble time will be the only available time scale in this situation. In fact, the solution in this case may be written as follows. $$ p_{conf} (t) = p_0 \left( 1 - H(t_0) \int^t_{t_0} d t' \exp \left( - \int^{t'}_{t_0} H(x) dx \right) \right) \exp \left( \int^t_{t_0} H(t') dt' \right) $$ where we have taken the following initial conditions; $$ p_{conf} (t=t_0) = p_0, \; \; \; {{dp_{conf}}\over {dt}}(t=t_0) = 0 $$ The second condition expresses the fact that the reaction force is absent at the initial time. The solution shows that the momentum decays in the Hubble time. Thus this process should not be simply neglected. However the above conclusion is obtained as a classical effect and it is not clear if the geometric bremsstrahlung is still effective when the quantum effect is taken into account. We shall discuss quantum geometrobremsstrahlung in the next section. \section{Quantum Geometric Bremsstrahlung} \hspace{0.8cm} As argued in the section 2, geometric bremsstrahlung may work efficiently classical mechanically in the early Universe. Whether this notable process survives or not after taking quantum effects into account is rather nontrivial and this question will be addressed next. To define well-behaved quantum amplitudes in the expanding Universe, we consider spacetimes with Minkowskian in- and out- regions. The way of expansion is chosen arbitrary. The scale factor is described as \begin{equation} a(\eta) = C(\lambda \eta)\label{1001} \end{equation} where $\eta$ is the conformal time, $\lambda^{-1}$ is a constant exhibiting typical time scale of the expansion. The function $C(x)$ in eqn(\ref{1001}) is arbitrary except the following constraints, \begin{eqnarray} C(x\sim -\infty ) &=&b ,\label{1011}\\ C(x\sim \infty) &=&1,\label{1012}\\ C(x) &>& 0,\label{1013} \end{eqnarray} where $b$ is some positive constant describing ratio of initial scale factor to final one. Consider first the photon emission process in massive scalar QED with conformal coupling to the background curvature. The action reads \begin{eqnarray} S= \int d^4 x &\sqrt{-g} &\left( (\nabla_{\mu} +ieA_{\mu}) \Phi^{\ast} (\nabla^{\mu} -ieA^{\mu} )\Phi \right. \nonumber\\ &&\left. +(\frac{1}{6} R-m^2 )\Phi^{\ast} \Phi - \frac{1}{4} F^{\mu\nu} F_{\mu\nu} \right).\nonumber \end{eqnarray} The photon emission process is prohibited in the flat spacetime by energy-momentum conservation. However in the expanding spacetimes the energy conservation law gets broken and the transition can take place. The transition amplitude is given in the lowest order of pertubation such that \begin{equation} Amp = -ie \int d^4 x \sqrt{-g} i\left( \Phi_f^{\ast}\nabla^{\mu} \Phi_i -\nabla^{\mu} \Phi_f^{\ast} \Phi_i \right) A^{\ast}_{\mu} \label{24} \end{equation} where $\Phi_i$($\Phi_f$) is initial(final) mode function of massive charged scalar field and $A^{\ast}_{\mu}$ is final mode function of electromagnetic field. The scalar mode functions satisfy \begin{equation} \left(\nabla^2 +m^2 -\frac{1}{6}R\right) \Phi =0.\nonumber \end{equation} Redefining the field $\tilde{\Phi}=\Phi\cdot a(\eta)$, the wave equation becomes the Klein-Gordon equation with a time-dependent mass, \begin{equation} \left(\partial^2 +m^2 a(\eta)^2 \right) \tilde{\Phi} =0.\label{1002} \end{equation} Here we introduce $g_{\vec{p}_i}^{in} (\eta)$ and $g_{\vec{p}_f}^{out} (\eta)$ satisfying a Schr{\"o}dinger-type equation, \begin{equation} \left[-\frac{d^2}{d\eta^2} -m^2 a(\eta)^2 \right]g_{\vec{p}} =\vec{p}^{\ 2} g_{\vec{p}},\label{9} \end{equation} with the boundary conditions in the asymptotic in and out regions as \begin{eqnarray} g_{\vec{p}_i}^{in} (\eta) &\rightarrow & \frac{ \exp \left( -i\eta \sqrt{\vec{p}_i^{\ 2} + m^2 b^2} \right) }{\sqrt{ (2\pi)^3 2 \sqrt{\vec{p}_i^{\ 2} +m^2 b^2}} } \ \ \ (\eta\sim -\infty),\nonumber\\ g_{\vec{p}_f}^{out} (\eta) &\rightarrow &\frac{\exp\left( -i\eta\sqrt{\vec{p}_f^{\ 2} +m^2} \right) }{ \sqrt{(2\pi)^3 2\sqrt{\vec{p}_f^{\ 2} +m^2 }} } \ \ \ (\eta\sim \infty).\nonumber \end{eqnarray} They also satisfy the following normalization condition. \begin{equation} i\left(g^{\ast}_{\vec{p}} {g'}_{\vec{p}} -{g'}^{\ast}_{\vec{p}} g_{\vec{p}} \right)=\frac{1}{(2\pi)^3 },\nonumber \end{equation} where the prime denotes the derivative with respect to $\eta$. Then the mode functions can be expressed as \begin{eqnarray} \Phi_f^{\ast} &=&\frac{1}{a}\tilde{\Phi}^{\ast}_f =\frac{1}{a} e^{-i\vec{p}_f \cdot \vec{x} }g^{out\ast}_{\vec{p}_f},\nonumber\\ \Phi_i &=&\frac{1}{a}\tilde{\Phi}_i =\frac{1}{a} e^{i\vec{p}_i \cdot \vec{x} }g_{\vec{p}_I}^{in}.\nonumber \end{eqnarray} The electromagnetic final mode function satisfies the Maxwell equation in curved spacetime, \begin{equation} \nabla^{\mu}( \nabla_{\mu} A^{\ast}_{\nu} -\nabla_{\nu} A^{\ast}_{\mu} ) =0.\label{23} \end{equation} Notice that in 4-dimensional conformally flat spacetimes this eqn(\ref{23}) can be reduced into the same form in the flat spacetime, \begin{equation} \partial^{\mu}( \partial_{\mu} A^{\ast}_{\nu} -\partial_{\nu} A^{\ast}_{\mu} ) =0.\nonumber \end{equation} Therefore we get easily the final mode function \begin{equation} A^{\ast}_{\mu} = \epsilon^{\ast}_{\mu} (\vec{k}) \frac{\exp\left( i|\vec{k} |\eta-i \vec{k}\cdot \vec{x} \right) }{ \sqrt{(2\pi)^3 2|\vec{k}|} },\nonumber \end{equation} where $\epsilon^{\ast}_{\mu}$ is a helicity factor. Using the rescaled field, the amplitude, eqn(\ref{24}), is rewritten as \begin{equation} Amp = -ie \int d^4 x i\left( \tilde{\Phi}_f^{\ast}\partial^{\mu} \tilde{\Phi}_i -\partial^{\mu} \tilde{\Phi}_f^{\ast} \tilde{\Phi}_i \right) A^{\ast}_{\mu}.\nonumber \end{equation} Because the photon emission lasts only during the epoch of expansion, the concept of the probability {\it per unit time} is ambiguious. So we shall use the transition probability itself. The transition probability $W$ can be obtained from the amplitude such that \begin{equation} W = \sum_{h=L, R} \frac{(2\pi)^3}{V}\int d^3 p_f d^3 k |Amp|^2,\nonumber \end{equation} where the summation is performed over the photon helicity and $V$ is the conformal volume of the space, which is cancelled by the factor $(2\pi)^3 \delta(\vec{0})$ coming from the conformal momentum conservation in $|Amp|^2$ . After the helicity summation, the explicit form of $W$ is obtained as follows. \begin{eqnarray} W &=& (2\pi)^6 e^2 \int \frac{d^3 k}{(2\pi)^3 2|\vec{k}|} \int d^3p_f \delta(\vec{k}+\vec{p}_f -\vec{p}_i ) \nonumber\\ &&\times\left[ (\vec{p}_f +\vec{p}_i )^2 \left| \int d\eta \ e^{i|\vec{k}|\eta} g^{out\ast}_{\vec{p}_f} g^{in}_{\vec{p}_i} \right|^2 -\left| \int d\eta\ e^{i|\vec{k}|\eta} ( g^{out\ast}_{\vec{p}_f} {g^{in}_{\vec{p}_i}}' -{g^{out\ast}_{\vec{p}_f}}' g^{in}_{\vec{p}_i} ) \right|^2 \right]\ .\nonumber \end{eqnarray} By virtue of the Wronskian relation; \begin{equation} \frac{d}{d\eta}\left( g^{out\ast}_{\vec{p}_f} {g^{in}_{\vec{p}_i}}' -{g^{out\ast}_{\vec{p}_f}}' g^{in}_{\vec{p}_i} \right) = (\vec{p}_f^{\ 2} -\vec{p}_i^{\ 2} ) g^{out\ast}_{\vec{p}_f} g^{in}_{\vec{p}_i}, \nonumber \end{equation} the form of $W$ is more simplified such that \begin{eqnarray} W= (2\pi)^3 e^2 \int \frac{d^3 k}{ 2|\vec{k}|} 4\left( \vec{p_i}^2 -\frac{ (\vec{k}\vec{p}_i )^2 }{\vec{k}^2 } \right) \left| \int^{\infty}_{-\infty} d\eta \ e^{i|\vec{k}|\eta} g^{out\ast}_{\ \vec{p}_i -\vec{k}} g^{in}_{\vec{p}_i} \right|^2 .\label{100} \end{eqnarray} To grasp the behavior of $W$ in the $|\vec{p}_i|\rightarrow \infty$ limit, we first argue a case with scale factor \begin{equation} a(\eta) = \Theta (\eta) + b \Theta (-\eta).\label{8} \end{equation} Then the exact mode functions are derived as \begin{eqnarray} g^{in}_{\vec{p}_i} &=& \Theta(-\eta) \frac{ \exp \left( -i\eta \sqrt{\vec{p}_i^{\ 2} + m^2 b^2} \right) }{\sqrt{ (2\pi)^3 2 \sqrt{\vec{p}_i^{\ 2} +m^2 b^2}} } \nonumber\\ &&+\Theta(\eta) \frac{ A(\vec{p}_i)\exp \left( -i\eta \sqrt{\vec{p}_i^{\ 2} + m^2} \right) +B(\vec{p}_i)\exp \left( i\eta \sqrt{\vec{p}_i^{\ 2} + m^2 } \right) }{\sqrt{ (2\pi)^3 2 \sqrt{\vec{p}_i^{\ 2} +m^2 b^2}} },\label{1}\\ g^{out}_{\vec{p}_f} &=& \Theta(\eta) \frac{ \exp \left( -i\eta \sqrt{\vec{p}_f^{\ 2} + m^2} \right) }{\sqrt{ (2\pi)^3 2 \sqrt{\vec{p}_f^{\ 2} +m^2 }} } \nonumber\\ &&+\Theta(-\eta) \frac{ A(\vec{p}_f)\exp \left( -i\eta \sqrt{\vec{p}_f^{\ 2} + m^2 b^2} \right) +B(\vec{p}_f)\exp \left( i\eta \sqrt{\vec{p}_f^{\ 2} + m^2 b^2} \right) }{\sqrt{ (2\pi)^3 2 \sqrt{\vec{p}_f^{\ 2} +m^2 }} },\nonumber\\ &&\ \label{2} \end{eqnarray} where \begin{eqnarray} A(\vec{p}) &=& \frac{1}{2}\left(1+\sqrt{ \frac{\vec{p}^{\ 2} +m^2 b^2} {\vec{p}^{\ 2} + m^2} }\right),\nonumber\\ B(\vec{p}) &=& \frac{1}{2}\left(1-\sqrt{ \frac{\vec{p}^{\ 2} +m^2 b^2} {\vec{p}^{\ 2} + m^2} }\right).\nonumber \end{eqnarray} Substituting eqn(\ref{1}) and eqn(\ref{2}) into eqn(\ref{100}) and taking the high momentum limit, $|\vec{p}_i|\rightarrow \infty$, it is shown that the terms proportinal to $B$ do not contribute to $W$ because of the damping behavior of $B$. Notice that taking the high momentum limit, the energy conservation law almost restores in the following sense. \begin{equation} |\vec{p}_i|\sim |\vec{k}|+|\vec{p}_i -\vec{k}|. \label{103} \end{equation} This can be read easily from the $\eta$ integration part in eqn (\ref{100}). Contribution from the region where eqn(\ref{103}) does not hold is severely suppressed by the energy conservation factor. Using the polar coordinate decomposition $\vec{p}_i \cdot \vec{k} =|\vec{p}_i |k \cos \theta$ with $k=|\vec{k}|$ and taking $|\vec{p}_i|$ much larger than $m$, it is easily derived that no contribution comes from the $k$ integral region lying between $|\vec{p}_i|$and $\infty$. Hence we get \begin{eqnarray} &&W(|\vec{p}_i|\sim \infty)\nonumber\\ &&= \frac{e^2 |\vec{p}_i|} {2 (2\pi)^2} \int^{|\vec{p}_i|}_{|\vec{p}_i| \delta} dk \frac{k}{|\vec{p}_i| -k} \int^{\pi}_0 d\theta\sin^3 \theta \nonumber\\ &&\times \left| \left[k-\sqrt{\vec{p}_i^{\ 2} +m^2 } +\sqrt{(|\vec{p}_i| -k)^2 +2|\vec{p}_i|k(1-\cos\theta) +m^2 } \right]^{-1} \right.\nonumber\\ &&\left. -\left[k-\sqrt{\vec{p}_i^{\ 2} +m^2 b^2 } +\sqrt{(|\vec{p}_i| -k)^2 +2|\vec{p}_i|k(1-\cos\theta) +m^2 b^2} \right]^{-1} \right|^2 .\nonumber\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \label{1112} \end{eqnarray} We need infra-red cutoff $|\vec{p}_i|\delta$ in eqn(\ref{1112}) due to the existance of massless photon. This infra-red divergence is well known one in flat spacetime quantum field theories with massless particles and it should be cancelled by an infra-red divergence of the self energy term \cite{BN}. The cutoff $\delta$ is physically determined by resolving power of soft photon observation. The $\theta$ integration in eqn(\ref{1112}) can be straightforwardly calculated. After performing this integration and taking the high momentum limit $|\vec{p}_i|\rightarrow\infty$, the $k$ integration is simplified and we finally obtain \begin{eqnarray} W(|\vec{p}_i|\rightarrow \infty) = \frac{e^2}{4\pi^2} \left( \ln \frac{1}{\delta}+\delta -1 \right) \left(\frac{1+b^2}{1-b^2}\ln \frac{1}{b^2} -2\right). \label{56} \end{eqnarray} Note that we have taken the helicity sum in eqn(\ref{56}). Instead, it is also possible to evaluate $W(b)$ independently with a fixed photon helicity. For left and right handed helicity, each probability is the same, a half of $W$ in eqn(\ref{56}). Furthermore we can also obtain the analytic forms of $W(b)$ in the spinor QED for the case of eqn(\ref{8}). Because both of the charged fermion and photon have degree of helicity freedom, 4 helicity contributions must be considered separately. The probability in the high momentum limit for 1/2 helicity fermion decaying into fermion with helicitity $h_{fermion}$ and photon with helicity $h_{photon}$ is denoted by $W(1/2; h_{fermion} , h_{photon} )$ and is given for each case as follows. \begin{eqnarray} &&W(1/2 ; 1/2 ,1) = \frac{e^2}{8\pi^2} \left( \ln \frac{1}{\delta} \right) \left(\frac{1+b^2}{1-b^2}\ln \frac{1}{b^2} -2\right).\label{12}\\ &&W(1/2 ; 1/2, -1) = \frac{e^2}{8\pi^2} \left( \ln \frac{1}{\delta} -\frac{\delta^2}{2} +2 \delta -\frac{3}{2} \right) \left(\frac{1+b^2}{1-b^2}\ln \frac{1}{b^2} -2\right).\label{13}\\ &&W(1/2;-1/2,1) = \frac{e^2}{8\pi^2} \left(1- \frac{b}{1-b^2}\ln \frac{1}{b^2} \right).\label{10}\\ &&W(1/2;-1/2, -1) = 0.\label{11} \end{eqnarray} No infra-red cutoff $\delta$ appears in eqn(\ref{10}) and eqn(\ref{11}) because spinflip of the fermion enables observers to distinguish the bremsstrahlung from the self energy process. \bigskip There exists a very useful aspect of $W$ in the high momentum limit. It is supposed that the results eqn(\ref{56})$\sim$eqn(\ref{11}) are exact not only for the special way of the expansion given by eqn(\ref{8}) but also arbitrary way satisfying eqn(\ref{1001}) $\sim$ eqn(\ref{1013}). This implies that $W(|\vec{p}_i|\rightarrow \infty)$ possesses a remarkable universality with respect to the ways of the cosmic expansion. This property may be explained as the Lorentz contraction effect from the view point of the high energy particle. Imagine a particle running in the comoving frame. Suppose that the Universe begins to expand when the particle passes through a point A and the Universe ceases to expand when the particle reaches point B. The particle catches energy from the expansion only while running from A to B. Taking the high momentum limit, the length between A and B contracts to zero in the rest frame of the particle. Therefore the particle cannot see the details of the way how the Universe expands and thus the universality of $W$ crops up. To see the universality more quantitatively, we shall discuss the scalar QED with an adiabatically slow evolution of the scale factor $a(\eta)$ satisfying eqn(\ref{1001}) $\sim$ eqn(\ref{1013}). In the zeroth order adiabatic approximation(WKB approximation) the mode functions satisfying eqn(\ref{9}) is written as \begin{eqnarray} g^{in}_{\vec{p}} \sim g^{out}_{\vec{p}} \sim \frac{ \exp \left[ i\vec{p}\cdot\vec{x} -i\int^\eta_0 d\eta' \sqrt{\vec{p}^{\ 2}+ m^2 a(\eta ')^2 } \right] } { \sqrt{ (2\pi)^3 2\sqrt{\vec{p}^{\ 2} +m^2 a(\eta )^2} } }.\label{101} \end{eqnarray} Substituting eqn(\ref{101}) into eqn(\ref{100}) and introducing the polar coordinate decomposition; $\vec{p}_i \cdot \vec{k} =|\vec{p}_i |k \cos \theta$, we get \begin{eqnarray} &&W \sim \frac{e^2 |\vec{p}_i|^2} {2 (2\pi)^2} \int^{\infty}_{|\vec{p}_i| \delta} dk k \int^{\pi}_0 d\theta\sin^3 \theta \nonumber\\ &&\times \left| \int^{\infty}_{-\infty} d\eta \ \frac{ \exp \left[ik\eta -i\int^\eta d\eta '\sqrt{\vec{p}_i^{\ 2} +m^2 a(\eta ')^2} +i\int^\eta d\eta '\omega(\vec{p}_i -\vec{k},\eta ') \right] } { \sqrt{\omega(\vec{p}_i -\vec{k},\eta ) \sqrt{\vec{p}_i^{\ 2} +m^2 a(\eta )^2} } } \right|^2\nonumber\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \label{102} \end{eqnarray} where $\omega(\vec{p}_i -\vec{k},\eta )= \sqrt{(|\vec{p}_i| -k)^2 +2|\vec{p}_i|k(1-\cos\theta) +m^2 a^2}$ and $|\vec{p}_i|\delta$ is the inra-red cutoff. Consider the high momentum limit in eqn(\ref{102}). As mentioned before nonvanishing contribution to $W$ comes from the integral region where the momentum holds the relation eqn(\ref{103}). Thus it is enough to restrict the momentum region between $|\vec{p}_i|$ and $|\vec{p}_i|\delta$. Here it should be searched which integral region of $\theta$ contributes, accompanied with the influence of $a(\eta)$, to the nonvanishing value of $W$ in eqn(\ref{102}). Due to eqn(\ref{103}), only emittion to nearly forward direction ($\theta\sim0$) is permitted and especially the integral region of $\theta$ safisfying $$0\leq \theta\leq O\left(m/|\vec{p}_i|\right)$$ gives the scale factor dependence to the $W$. Several expansions like \begin{eqnarray} \sqrt{\vec{p}_i^{\ 2} +m^2 a(\eta ')^2} \sim |\vec{p}_i |+\frac{m^2 a(\eta ')^2}{2|\vec{p}_i |} \nonumber \end{eqnarray} yield finally \begin{eqnarray} &&W \sim \frac{e^2 |\vec{p}_i|}{2 (2\pi)^2} \int^{|\vec{p}_i|}_{|\vec{p}_i|\delta} dk \frac{k}{|\vec{p}_i| -k} \int^{O\left(\frac{m}{|\vec{p}_i|}\right) }_0 d\theta\ \theta^3 \nonumber\\ &&\times \left| \int^{\infty}_{-\infty} d\eta \ \exp \left[ -i\int^\eta_0 d\eta ' \left( \frac{m^2 a(\eta ')^2 }{2|\vec{p}_i|} -\frac{m^2 a(\eta ')^2}{2(|\vec{p}_i| -k)} -\frac{|\vec{p}_i|k\theta^2 )}{2(|\vec{p}_i| -k)} \right) \right] \right|^2 \nonumber\\ &&= \frac{ e^2 }{(2\pi)^2} \int^1_{\delta} dy \frac{y}{1-y} \int^{O(1)}_0 dz z^3 \nonumber\\ &&\times \left| \int^{\infty}_{-\infty} d\tilde{\eta} \ \exp \left[ -i\int^{\tilde{\eta}}_0 d\tilde{\eta}' \left( C\left(\frac{2\lambda|\vec{p}_i|}{m^2}\tilde{\eta}'\right)^2 -\frac{C(\frac{2\lambda|\vec{p}_i|}{m^2}\tilde{\eta}')^2 }{1-y} -\frac{y z^2}{1-y} \right) \right] \right|^2 ,\nonumber\\ &&\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \label{104} \end{eqnarray} where we change the integral variables in the following way, \begin{eqnarray} k&=& |\vec{p}_i| y,\nonumber\\ \theta&=& \frac{m}{|\vec{p}_i|}z,\nonumber\\ \eta&=& \frac{2|\vec{p}_i|}{m^2} \tilde{\eta}.\nonumber \end{eqnarray} Note that the function $C(\frac{2\lambda|\vec{p}_i|}{m^2}\tilde{\eta})$ in eqn(\ref{104}) approaches in the high momentum limit to a step function, \begin{eqnarray} C\left(\frac{2\lambda|\vec{p}_i|}{m^2}\tilde{\eta}\right) \sim \Theta (\tilde{\eta}) + b\ \Theta(-\tilde{\eta}).\label{1.11} \end{eqnarray} Therefore the value of $W$ for arbitrary adiabatical cosmic expansion satisfying eqn(\ref{1001}) $\sim$ eqn(\ref{1013}) must equal to the specified value for eqn(\ref{8}), and the universality is surely realized. If we dismiss the adiabatic approximation, the mode functions have reflection wave terms like in eqn(\ref{1}) and eqn(\ref{2}). However amplitude of the reflection waves vanishes in the high momentum limit and the universality are thought to survive. \bigskip Here we have a comment on the rate of the geometric bremsstrahlung. Comparing with the classical results in the section 2, it is noticed from eqn(\ref{56})$\sim$ eqn(\ref{11}) that interaction rate of quantum geometric bremsstrahlung does not so large compared with classical one. This is due to the fact that quantum effect smears position of the classical point particle and dilutes the charge density. \bigskip Now it is worth considering implication of the results; eqn(\ref{56})$\sim$eqn(\ref{11}) in connection with the conformal symmetry. Since the massless limit $m\rightarrow 0$ forces the speed of the particle to reach the light velocity, the universality with respect to the way of the cosmic expansion is maintained. In the lowest pertubation of the QED, the massless limit is shown to be equivalent with the $|\vec{p}_i| \rightarrow \infty$ limit. Therefore again the same results, eqn(\ref{56})$\sim$ eqn(\ref{11}), come up in the $m\rightarrow 0$ limit and $W$ really possesses the non-vanishing value. One might naively expect for the massless case that the amplitude in the conformally flat spacetime vanishes as in the Minkowskian spacetime, by virtue of the conformal symmetry guaranteed at least classical mechanically. However this is {\it not} true unless $b=1$ as argued above. \bigskip Next let us discuss the rate of the geometric bremsstrahlung. Taking large expansion limit $b\sim 0$, the transition probablity can be typically expressed as \begin{eqnarray} W(b\sim 0) \sim O(1) \frac{N^{\ast}e^2}{4\pi^2} \ln \frac{1}{b } , \end{eqnarray} where $N^{\ast}$ is the number of final modes. In our spinor QED model we take $N^{\ast} =2$ due to contributions of eqns (\ref{12}) and (\ref{13}). However in the standard model we have more particles and $N^{\ast} \sim O(10)$. Moreover extended theories (like GUT and SUSY) can gives us $N^{\ast} \sim O(100)$. The process is thought to occur when $W\sim 1$. Thus when the Universe expands enough to satisfy \begin{eqnarray} \frac{1}{b}=\frac{a_f}{a_i}\sim \exp \left[ \frac{4\pi^2 O(1)}{N^{\ast} e^2} \right] \label{555} \end{eqnarray} the event will take place. If we specify a model of the Unvierse evolution, we can get the rate itself. For example let us assume that the expansion is dominated by the radiation, $a(t) \propto t^{1/2}$. Then we get the following estimation for the transition rate. \begin{eqnarray} \Gamma \sim \frac{1}{t_f} = b^2 \frac{1}{t_i} \sim \exp\left[ - \frac{8\pi^2 O(1)}{N^{\ast} e^2} \right] H_i \sim e^{-\frac{O(100)}{N^{\ast}} }H_i, \end{eqnarray} where $H_i =1/t_i$ is the Hubble parameter at the initial time. Thus we cannot neglect the quantum bremsstrahlung naively when $N^{\ast}\sim O(100)$. It might be also worth reminding that rates of ohther gauge interection processes different from the geometric bremsstrahlung is much smaller than the Hubble parameter when the temparature of the Universe is higher than $O(10^{15})$ GeV. Thus the bremsstrahlung may be the most dominant process in such a early era. \section*{ Acknowledgement} The authors wish to thank T. Goto, I.Joichi, T. Moroi, M. Tanaka and M. Yoshimura for fruitful discussions. We also thank to K.Hikasa and J.Arafune for their critical comments.
\section{Motivation} As a result of work in the 1960's and 1970's \cite{Allan}, some of which has continued beyond then, it is recognized that air showers of energy 10$^{17}$ eV are accompanied by radio-frequency pulses, whose polarization and frequency spectrum suggest that they are due mainly to the separation of positive and negative charges of the shower in the Earth's magnetic field. The most convincing data have been accumulated in the 50--100 MHz frequency range. However, opinions have differed regarding the strength of the pulses, and atmospheric and ionospheric effects have led to irreproducibility of results. In particular, there may also be pulses associated with cosmic-ray-induced atmospheric discharges \cite{atm}. A study is being undertaken of the feasibility of equipping the Auger array with the ability to detect such pulses. It is possible that the higher energy of the showers to which the array would be sensitive would change the parameters of detection. Before a design for large-scale RF pulse detection can be produced, it is necessary to retrace some of the steps of the past 30 years by conclusively demonstrating the existence of the pulses for 10$^{17}$ eV showers, and by controlling or monitoring some of the factors which led to their irreproducibility in the past. In this note we describe the prototype activity at the CASA/MIA site, mention related activities, and set forth some considerations regarding plans for the Auger project. More concrete plans for RF detection must await the outcome of protoyping work at the CASA/MIA site. \section{CASA/MIA Prototype setup} In order to verify the claim \cite{Allan} that 10$^{17}$ eV showers are accompanied by RF pulses with significant energy in the 50--100 MHz range, a prototype detector is being set up at the CASA/MIA site in Dugway, Utah. This section describes the status of that effort. \subsection{Large-event trigger} A trigger based on the coincidence of several muon ``patches'' was set to select ``large'' showers with a rate of several per hour. The MIA Patch-Sum trigger was sent into a fan-in/fan-out. From this, the signal was put into a LeCroy 821 Discriminator. This produced a pulse of height $-0.8$ V (on 50 Ohm output). The width of the output pulse was set to 200 nsec. The frequency of the output could be varied by adjusting the threshold on the discriminator. The observed rates as a function of this threshold were as follows: \begin{center} \begin{tabular}{||c|c|c||} \hline Threshold & Patches & Rate \\ \hline $-350$ mV & 6 & 45 Hz (highly variable) \\ $-400$ mV & 7 & $3.7 \pm 0.25$ Hz \\ $-450$ mV & 8 & $2.21 \pm 0.19$ Hz \\ $-500$ mV & 9 & $1.23 \pm 0.10$ Hz \\ $-560$ mV & 10 & $1.00 \pm 0.07$ Hz \\ $-690$ mV & 12 & $0.52 \pm 0.03$ Hz \\ $-750$ mV & 13 & $ 5.8 \pm 0.8 $ min$^{-1}$ \\ $-810$ mV & 14 & $0.65 \pm 0.18$ min$^{-1}$ \\ $-870$ mV & 15 & $ 2.0 \pm 0.8$ hr$^{-1}$ \\ \hline \end{tabular} \end{center} The rate varied somewhat depending on how noisy any particular patch is. In particular, below 6 patches, the rate was not observed to be very stable. The electronics was initially set up at the 15 patch level. This was estimated to be $5 \times 10^{16}$ eV to $10^{17}$ eV, based on the rate at $10^{18}$ eV of 1/km$^2$/day/sr. The performance of this trigger was monitored during two runs (17985 and 17987) on June 30, 1995. During a period in which 26 million CASA triggers had been registered (estimated to be about 16 days assuming a 20 Hz rate), 240 ``large-event triggers'' had been registered, or about 15 per day. The counters were reset on the evening of 6/29/95. During the subsequent 12.5 hours, 7 counts were registered between 7:12 pm MDT 6/29 and 7:47 am MDT 6/30, consistent with the above rate. This is considerably less than the hoped-for several per hour or the 2 per hour noted in the table above. On the morning of 6/30/95, the large-event trigger level was reset from $-869$ mV to $-822$ mV. At this level, it was checked that several counts per hour would be registered. The large-event trigger was used to key a transceiver which then broadcast a digitally recorded voice message which was picked up remotely. Times of receipt of the trigger message were measured to an accuracy of about 5 seconds. During remote receipt of trigger notifications, other tasks were being performed, such as setting up of antenna, monitoring of RF backgrounds, and testing of preamplifier. Thus it is possible that some notifications were missed. For future studies it would be very helpful to record ``large-event triggers'' with a time stamp and, if possible, a pulse voltage level to determine by what amount a given threshold was exceeded. A file of events with at least 15 muon patches was generated after the day's runs: \begin{center} \begin{tabular}{c c c c} \hline Run & Events & Live time & Events/minute \\ & & (minutes) & \\ \hline 17985 & 446 & 354 & $1.26 \pm 0.06$ \\ 17987 & 495 & 354 & $1.40 \pm 0.06$ \\ Both & 941 & 708 & $1.33 \pm 0.04$ \\ \hline \end{tabular} \end{center} The times of events in this file which matched within 5 seconds of those found using the remote-monitoring system mentioned were recorded. Of 58 potential matches between large-event triggers and events in the above file, only 27 actual matches were found. Although this correlation appears higher than accidental, it is clear that many large-event triggers failed to match with events in the actual data record. Consequently, more effort is being devoted to construction of an efficient large-event trigger. It is notable that in measuring HiRes - MIA coincidences for events of 10$^{17}$ eV, a rate of somewhat less than 1 per hour was achieved \cite{coincs}. \subsection{Monitoring of RF noise environment} It was a concern that the RF noise of the local electronics and the presence of an extensive lightning-protection array might dictate the placement of one or more antennas outside the periphery of the array, or might make the site unsuitable altogether. The behavior of a single CASA board was investigated at the University of Chicago. The various clock signals were detected at short distances ($< 1$ m) from the board, but a much more intense set of harmonics of 78 kHz emanated from the switching power supplies. These harmonics persisted well above 100 MHz. At 144--148 MHz, they overlapped, leading to intense broad-band noise. An initial survey of RF noise at the CASA site was performed. On the basis of the results, which indicated some RF noise within the array, it was decided to perform an initial follow-up survey sitting just outside the array. The original log-periodic antenna used to detect RF pulses at Chacaltaya in the 1960's and 1970's was obtained, tested for bandwidth, taken out to Utah, and used in a follow-up study of RF noise in the 60--85 MHz frequency range. Sources of most strong RF signals in this range appeared to be due to either the receiver itself or to local TV stations. Spectrum analysis techniques may be suitable for removing such monochromatic signals. \subsection{Near-term plans} It is proposed to monitor the RF noise and to detect pulses by mounting the log-periodic antenna near the CASA central trailer site, just above the lightning protection grid. A digital storage scope will be used to register several microseconds of RF data on a rolling basis. These data will then be captured and inspected visually upon receipt of a large-event trigger. The experiment will be repeated using successively greater amounts of amplification and narrower band-pass filters once these become available. The filters are being developed at the University of Chicago. Once the large-event trigger has been demonstrated to select events of 10$^{17}$ eV and above, permanent digital recording of coincident pulses will be undertaken. Still to be performed are experiments which seek to monitor RF pulses at lower frequencies and at greater distances from the array. For these pulses, whose strengths may be correlated with atmospheric electric fields, it is planned to monitor such fields with the help of a field mill. A spectrum analyzer will be used to make a broader survey of the RF noise in various frequency ranges and may be of help in detecting potential sources of interference to RF communications in the Auger project. \section{Recent information on related activities} \subsection{FORTE, BLACKBEARD, SNO, and other projects requiring digitizers} Discussions with John Wilkerson at the University of Washington have been very productive. Wilkerson was engaged in projects at Los Alamos with the acronyms FORTE and BLACKBEARD whose aim was detection of electromagnetic pulses, including those produced by cosmic-ray-induced electromagnetic discharges, with frequency ranges in the 30--100 MHz range. Many of the fast-digitization and memory problems appear to be identical to those in the proposal for a protoype pulse detector at CASA/MIA. Time-frequency plots have been obtained by the BLACKBEARD project which are exactly those one would hope to generate in a survey at CASA/MIA. Wilkerson has also encountered requirements similar to ours for digitization of SNO data. His estimate is that one can use Maxim MAX 100 A/D chips for less than \$1K per channel, but that feeding their output into memory may well amount to another \$1K per channel. Other references on digitizers have been obtained \cite{Atiya,Bryman}. Discussions with Wilkerson will continue, and further discussions with Dan Holden at Los Alamos are envisioned. \subsection{Status of GHz detection} David Wilkinson, who visited the University of Chicago during the spring of 1995, has promised to look into the power radiated at frequencies of several GHz, where new opportunities exist associated with the availability of low-noise receivers. At latest report he had planned to complete the relevant calculations during the summer of 1995. \subsection{Other options} Dispersion between arrival times of GPS signals on two different frequencies may serve as a useful monitor of air shower activity. The possibility of correlation of large showers with such dispersion events will be investigated. It may be possible at the CASA/MIA site to monitor commercial broadcast signals in the 55 - 88 MHz range to detect momentary enhancements associated with large showers, in the same sense that meteor showers produce such enhancements. Television Channels 3 and 6, for which no nearby stations exist, offer one possibility. \section{Considerations for Auger project} At present we can only present a rough sketch of criteria for detection in the 50--100 MHz range. Data would be digitized at a 500 MHz rate at each station and stored in a rolling manner, with at least 10 microseconds of data in the pipeline at any moment. Upon receipt of a trigger signaling the presence of a ``large'' shower ($> 10^{17}$ eV), these data would be merged into the rest of the data stream at each station. Per station, we estimate the following additional costs, in US dollars, for RF pulse detection: \begin{center} \begin{tabular}{l l c} \hline Two antennas and impedance transformers: & 200 & (a) \\ Mounting hardware: & 100 & (b) \\ Cables and connectors: & 200 & (c) \\ Preamps and lightning protection: & 100 & (d) \\ Digitization and memory electronics: &2000 & (e) \\ \hline Total per station: &2600 & (f) \\ \hline \end{tabular} \end{center} \noindent \noindent (a) Two commercial log-periodic TV antennas with commercial 4:1 baluns; crossed polarizations. Difference signal to be detected. \noindent (b) Highly dependent on other installations at site. Antennas are to be pointed vertically but optimum elevation not yet determined. \noindent (c) Antennas are mounted near central data acquisition site of each station. \noindent (d) Commercial GaAsFET preamps and gas discharge tubes. \noindent (e) Subject to prototype development experience. Power requirements not yet known. \noindent (f) The number of stations equipped with RF detection will not exceed 2000 per array, but could easily be fewer, depending on prototype experience. \bigskip The above estimate assumes that one can power the preamps and DAQ electronics from the supply at each station without substantial added cost. It also assumes that the ``large-event trigger'' will be available at each station. A further assumption is that the difference signal suffices to characterize the pulse. Additional preamplification and DAQ electronics may be required if this is not so. A major consideration may be the acquisition of antennas robust enough to withstand extreme weather (particularly wind) conditions. For detection at frequencies above or below 50--100 MHz, the criteria are not yet well enough developed to permit any cost estimate. \section{Acknowledgments} I thank Jim Cronin for inviting me to consider these questions for the Auger project, Lucy Fortson for help in surveying RF noise generated by a CASA board, Kevin Green for setting up the ``large-event trigger,'' Dick Gustafson for discussions and for information about SSC equipment, Gerard Jendraszkiewicz and Dave Smith for technical advice in the design of the ``alert module,'' Larry Jones for supplying the original antennas used on Mount Chacaltaya and for discussions, Brian Newport for logistical help at the CASA/MIA site and for generating the file of events from runs 17985 and 17987, Rene Ong for performing the initial RF survey work at the CASA/MIA site, Dave Peterson for the loan of equipment used for monitoring RF at the CASA/MIA site, Leslie Rosenberg for first interesting me in this problem, Fritz Toevs for technical advice and laboratory space at the University of Washington, Augustine M. Urbas for help with antenna measurements and filter design, John Wilkerson for discussions regarding electronics and RF pulse detection, Dave Wilkinson for stressing the importance of on-line monitoring of events during the initial RF survey work, and Bob Williams for laboratory space. This work was performed in part during a visit to the Institute for Nuclear Theory at the University of Washington, and was supported in part by the United States Department of Energy under Grant No.~DE FG02 90ER40560.
\section{Introduction} \label{int} The study of the electroweak interaction between leptons and hadrons has been a challenging topic ever since the standard model was proposed by Glashow, Salam and Weinberg. This model predicts the coupling of the electroweak currents to leptons and quarks in terms of the electric charge $e$ and the Weinberg angle $\theta_{W}$. In particular, the weak neutral current is mediated by the exchange of the $Z^0$ gauge boson. At low and moderate momentum transfer its contribution is suppressed relative to photon exchange by a factor $q^2/M_Z^2$, where $q$ is the four-momentum of the exchanged boson and $M_Z$ the mass of the $Z^0$. The interference term beween photon and $Z^0$ exchange contains a parity violating (PV) effect, which becomes visible as an asymmetry by scattering polarized electrons with helicity along the direction of the beam $h = +1$ or opposite to it $h = -1$, \begin{equation} \label{def_a} A = \frac{ \sigma(h = +1) - \sigma(h = -1)}{ \sigma(h = +1) + \sigma(h = -1)} \, . \end{equation} The quantity $\sigma$ in this equation should represent an inclusive cross section. In the case of a coincidence experiment, e.g. $e + N \rightarrow e + N + \pi$, there also appear parity conserving asymmetries due to the electromagnetic interaction. Specifically, the so-called 5th response function will generate a background of helicity-dependent contributions, which are parity conserving and, therefore, generally larger than the PV effects by several orders of magnitude. The pioneering experiment to measure PV asymmetries has been performed at SLAC \cite{Pr78} by deep inelastic electron scattering on a deuterium target. This experiment has been followed by investigations of PV quasifree scattering off $^9 B$ at Mainz \cite{He89} and PV elastic scattering off $^{12} C$ at MIT/Bates \cite{So90}. In the latter case the momentum transfer was only $q^2 = - (150 \mbox{ MeV})^2$, leading to the tiny asymmetry $A = (0.60\pm 0.14 \pm 0.02) \cdot 10^{-6}$. These experiments were originally devised to test the standard model, in particular to measure the Weinberg angle. The value obtained for this angle by high-energy experiments was confirmed within the error bars of about 10\%. With the advent of new electron accelerators like CEBAF, MAMI and MIT/Bates, having high intensity, high duty-factor and a polarized beam, the quality of the data can be considerably improved. Moreover, with $\sin^2\theta_W = 0.2319 (5)$ known to 4 decimal places and the standard model firmly established, the strategy of the new experiments will be redirected towards an improved understanding of the structure of the nucleon. In particular, PV elastic electron scattering will provide information on the strangeness content of the nucleon. Three such experiments are being planned with the 4 GeV CEBAF beam \cite{Be92,So91,Bei91} and there are also proposals to measure the electric radius and the magnetic moment of strange quark pairs at MAMI \cite{Ha93} and MIT/Bates \cite{MK89} in the region of 1 GeV. Assuming that the present experimental activities will soon yield novel information on the ground state of the nucleon, we deem it appropriate to study the effect of PV interactions for inelastic processes. It is therefore the aim of this contribution to investigate the PV asymmetries for electroproduction of pions. Previous calculations of Nath et al.\cite{Na82} and Jones et al.\cite{Jo80} at medium energies and of Cahn et al.\cite{Ca78} at higher energies have been based on the production of stable $\Delta$ isobars. However, there should be non-negligible background contributions interfering coherently with the resonance. Estimates for such contributions have been reported earlier by Ishankuliev et al.\cite{Is80} and by Li et al.\cite{Li82}. In particular, Li et al.\cite{Li82} have also considered PV effects in the hadronic wave function, i.e. at the $\pi N$ vertex. It turns out, however, that such contributions are of the order of $10^{-7}$, much smaller than the expected asymmetries due to PV interferences in the electroweak interaction, which are of the order of $10^{-4} q^2/\mbox{ GeV}^2$. In the following section we will briefly review the kinematics, the cross section and the decomposition of both vector and axial currents into the invariant amplitudes. Sect. 3 presents our model using effective Lagrangians. It includes a background of Born terms with pseudovector $\pi N$ coupling and the $\Delta$ isobar treated as a Rarita-Schwinger field with phenomenological $N\Delta$ transition currents. Assuming that the hadronic currents are dominated by $u$ and $d$ quarks, the weak neutral current may be decomposed in terms of strong isospin \cite{Mu94} and related to hadronic currents. The matrix elements are then decomposed into invariant amplitudes \cite{Ad68} and according to their isospin structure. The numerical results are presented in Sect. 4. As a test we have first calculated the inclusive electromagnetic cross section. The experimental data for this process \cite{Ly67,Co67} can be well reproduced by the model. We present our results for PV asymmetries as function of excitation energy, scattering angle and momentum and compare our asymmetries with previous calculations. Finally, we give a summary and some conclusions in Sect. 5. \section{Formalism} \label{form} \subsection{Kinematics} We consider the reaction shown in Fig. \ref{fig1} . $P_i = (E_i,\vec{P_i})$ and $P_f = (E_f,\vec{P_f})$ are the 4-momenta of the nucleon in the initial and the final state, respectively, the produced pion has momentum $k_\pi = ( \omega_\pi,\vec{k_\pi})$. The momentum transfer $q = ( \omega,\vec{q} )$ is the difference of the 4-momenta of the ingoing and outgoing electron $k_i - k_f$, with $k_l = ( \varepsilon_l,\vec{k_l})$. The spins of nucleons and electrons in the initial and final states are denoted by $S_i$, $S_f$, $s_i$, and $s_f$. If not stated otherwise, the kinematical variables are evaluated in the laboratory system, which is defined by $P_i = (m,\vec {0})$. Furthermore, the Mandelstam variables \begin{eqnarray} s &=& ( P_i + q )^2 ,\quad t = ( q - k_\pi)^2 , \quad u = ( P_i - k_\pi)^2 , \end{eqnarray} are equivalent to the following set of Lorentz invariant kinematic variables \cite{Ad68}: \begin{eqnarray} \nu &=& \frac{P \cdot q}{m} \, ,\qquad \nu_B = -\frac{ k_\pi \cdot q }{ 2 m} \, , \qquad W = \sqrt{s} \, , \end{eqnarray} with $P = \frac{1}{2}(P_i + P_f)$. The latter set of variables will be used from now on. Besides, the coincidence experiment is characterized by the angles shown in Fig. \ref{fig2} . The polar angle $\Theta_\pi$ is the angle between the pion 3-momentum $\vec{k}_\pi$ and the momentum transfer $\vec {q}$, the azimuthal angle $\phi_\pi$ is the angle between the scattering plane, defined by $\vec{k_i}$ and $\vec{k_f}$, and the reaction plane, spanned by $\vec{k_\pi}$ and $\vec{q}$. Furthermore, the scattering angle of the electron is $\Theta_e$. In the following sections we will be interested in the inclusive cross section, i.e. the cross section has to be integrated over the pion angles. \subsection{Invariant matrix element} The differential cross section \cite{Re90} \begin{equation} \label{wq} d \sigma = \frac{ (2 \pi)^4 \delta^4( P_i + q - P_f - k_\pi )}{4 [ (P_i \cdot k_i)^2 - m_e^2 m^2 ]^{\frac{1}{2}}} \prod^{n_f}_{j = 1} \frac{ d^3 \vec{p}_j }{ ( 2 \pi)^3 2 E_j } \mid {\cal M}_{fi} \mid^2 , \end{equation} has to be integrated over the momenta of the pion and the final nucleon. In the one boson exchange approximation and for momentum transfers $q^2 \ll {M_Z}^2$, the invariant matrix element is of the form \begin{eqnarray} \frac{q^4}{e^4} \mid {\cal M}_{fi} \mid^2 &=& \mid j_\nu^{EM} J^\nu_{EM} + \frac{q^2}{{M_Z}^2} j_\nu^{NC} J^\nu_{NC} \mid^2 + O( \frac{q^4}{{M_Z}^4} ) \\ & &= {j_\mu^{EM}}^\dagger j_\nu^{EM} {J^\mu_{EM}}^\dagger J^\nu_{EM} + \frac{q^2}{{M_Z}^2} \left( {j_\mu^{EM}}^\dagger j_\nu^{NC} {J^\mu_{EM}}^\dagger J^\nu_{NC} + h.c. \right) + O( \frac{q^4}{{M_Z}^4} ) \nonumber \\ & &= \eta_{\mu\nu}^{EM} W^{\mu\nu}_{EM} + \frac{q^2}{{M_Z}^2} \eta_{\mu\nu}^{I} W^{\mu\nu}_{I} + O( \frac{q^4}{{M_Z}^4} ) \; . \nonumber \end{eqnarray} In this expression, the photon couples to the electromagnetic current of the nucleon, $J_\nu^{EM} = V_\nu^{EM}$, and the $Z^0$ gauge boson to the familiar combination of vector and axial currents $J_\nu^{NC} = V_\nu^{NC} + A_\nu^{NC}$. The corresponding currents of the electron are denoted by $j_\nu^{EM}$ and $j_\nu^{NC}$. These currents may be combined to $ W^{\mu\nu}$ and $\eta_{\mu\nu}$, the hadronic and leptonic tensors. The tensors may be decomposed into symmetric and antisymmetric parts. If we neglect the electron mass, the leptonic tensor is \begin{eqnarray} \label{lept} \eta_{\mu\nu}^{(s)} &=& 2 ( 2 K_\mu K_\nu + \frac{1}{2} ( q^2 g_{\mu\nu} - q_\mu q_\nu ) ) \, ,\\ \eta_{\mu\nu}^{(a)} &=& - 2 i \epsilon_{\mu\nu\alpha\beta} q^\alpha K^\beta \, , \nonumber \end{eqnarray} with $q = ( k_i - k_f )$ and $K = \frac{1}{2} ( k_i + k_f )$. A reasonable definition for the hadronic tensor is \begin{eqnarray} \label{w} W_{\mu\nu}^{EM} &=& \frac{1}{2} \sum_{S_i,S_f} \langle P_f | \hat J_\mu^{EM}| P_i \rangle^\dagger \langle P_f | \hat J_\nu^{EM} | P_i \rangle \, , \\ W_{\mu\nu}^{I} &=& \frac{1}{2} \sum_{S_i,S_f} \left[ \langle P_f | \hat J_\mu^{EM} | P_i \rangle^\dagger \langle P_f | \hat J_\nu^{NC} | P_i \rangle + h.c. \right] \, . \nonumber \end{eqnarray} \subsection{Invariant amplitudes and isospin structure} The hadronic matrix elements have the structure \begin{equation} \label{stru} {\cal M} = \epsilon^\mu J_\mu \, , \end{equation} where $\epsilon^\mu$ is an abbreviation for the leptonic matrix element. These matrix elements may be decomposed in isospace according to \begin{equation} \label{iso} {\cal M} = \underbrace{\chi_{f}^{\dagger} \tau_\pi \chi^{\ }_{i} }_{=: I^0} {\cal M}^0 + \underbrace{\chi_{f}^{\dagger} \frac{1}{2} \{ \tau_\pi , \tau_3 \} \chi^{\ }_{i} }_{=: I^+} {\cal M}^+ + \underbrace{\chi_{f}^{\dagger} \frac{1}{2} [ \tau_\pi , \tau_3 ] \chi^{\ }_{i} }_{=: I^-} {\cal M}^- \, , \end{equation} with $\chi^{\ }_{f}$ and $\chi^{\ }_{i}$ the isospinors of the nucleon in the final and initial state, respectively, and $\tau_\pi$ the isospin matrix characterizing the produced pion. Furthermore, the hadronic transition currents are decomposed into invariant amplitudes \cite{Ad68} \begin{eqnarray} \label{ia} V_\mu^{(\pm, 0)} &=& \sum_{j = 1}^6 V_{j}^{(\pm, 0)}(\nu,\nu_B,q^2) \bar u(\vec{P_f}) M_\mu^j u(\vec{P_i}) \, , \\ A_{\mu}^{(\pm, 0)} &=& \sum_{j = 1}^8 A_{j}^{(\pm, 0)}(\nu,\nu_B,q^2) \bar u(\vec{P_f}) N_\mu^j u(\vec{P_i}) \nonumber \, . \end{eqnarray} The amplitudes $V_{j}^{(\pm, 0)},A_{j}^{(\pm, 0)}$ depend on the three independent variables $\nu,\nu_B$, and $q^2$, and the superscript $(\pm, 0)$ refers to the isospin decomposition (\ref{iso}). A reasonable choice for the vector $(M_\mu^j)$ and axial vector $(N_\mu^j)$ operators is \cite{Ad68}: \begin{equation} \begin{array}{l@{\qquad}l} M_\mu^1 = \frac{i}{2} \gamma_5 ( \gamma_\mu \dida{q} - \dida{q} \gamma_\mu ) & \zeta_1^V = 1 \\ M_\mu^2 = - 2 i \gamma_5 ( P_\mu \; k^\pi \cdot q - P \cdot q \; k^\pi_\mu ) & \zeta_2^V = 1 \\ M_\mu^3 = i \gamma_5 ( \gamma_\mu \; k^\pi \cdot q - \dida{q} k^\pi_\mu ) & \zeta_3^V = -1 \\ M_\mu^4 = 2 i \gamma_5 [ ( \gamma_\mu \, P \cdot q - \dida{q} P_\mu) -\frac{m}{2}(\gamma_\mu \dida{q} - \dida{q} \gamma_\mu ) ] & \zeta_4^V = 1 \\ M_\mu^5 = - i \gamma_5 ( q_\mu \; k^\pi \cdot q - q^2 k^\pi_\mu ) & \zeta_5^V = -1 \\ M_\mu^6 = i \gamma_5 ( q_\mu \dida{q} - q^2 \gamma_\mu ) & \zeta_6^V = -1 \label{ia_v} \end{array} \end{equation} and \begin{equation} \label{ia_a} \begin{array}{l@{\qquad}l} N_\mu^1 = \frac{i}{2} ( \gamma_\mu \dida{k^\pi} - \dida{k^\pi} \gamma_\mu ) & \zeta_1^A = -1 \\ N_\mu^2 = -2 i P_\mu & \zeta_2^A = -1 \\ N_\mu^3 = -i k^\pi_\mu & \zeta_3^A = 1 \\ N_\mu^4 = -i m \gamma_\mu & \zeta_4^A = -1 \\ N_\mu^5 = 2 i \dida{q} P_\mu & \zeta_5^A = 1 \\ N_\mu^6 = i \dida{q} k^\pi_\mu & \zeta_6^A = -1 \\ N_\mu^7 = -i {q}_\mu & \zeta_7^A = 1 \\ N_\mu^8 = i \dida{q} {q}_\mu & \zeta_8^A = -1. \\ \end{array} \end{equation} The vector current operators are explicitly gauge invariant by construction, $q \cdot M^j \equiv 0 \quad \forall j = 1 \ldots 6$. The constants $\zeta_j^V$ and $\zeta_j^A$ specify the behavior of the invariant amplitudes under the crossing transformation $\nu \rightarrow -\nu$\ \cite{Ad68}, \begin{eqnarray} \label{cross} V_{j}^{(\pm, 0)}(\nu,\nu_B,q^2) &=& (\pm, +) \zeta_j^V V_{j}^{(\pm, 0)}(-\nu,\nu_B,q^2) \, , \\ A_{j}^{(\pm, 0)}(\nu,\nu_B,q^2) &=& (\pm, +) \zeta_j^A A_{j}^{(\pm, 0)}(-\nu,\nu_B,q^2) \nonumber \, . \end{eqnarray} \subsection{Multipole decomposition} The vector and axial currents of (\ref{ia}) may be decomposed into a multipole series following the work of Adler \cite{Ad68}. The leading $S$-wave contributions to the currents are \begin{eqnarray} \label{mulzer} \epsilon \cdot V &=& \frac{4 \pi i W}{m} \chi_f^{\dagger} \big\{ \frac{\vec{\sigma} \cdot \hat{q}}{\mid \vec{q}\mid \omega} (\epsilon_0 \, Q^2 + \omega \; \epsilon \cdot q) L_{0+} - \vec{\sigma} \cdot \vec{\epsilon}_T E_{0+} + \ldots \big\} \chi_{i} \; , \\ \epsilon \cdot A &=& \frac{4 \pi W}{m} \chi_{f}^{\dagger} \big\{ \epsilon_0 \, {\cal L}_{0+} + i \vec{\sigma} \cdot (\hat{q} \times \vec{\epsilon}) {\cal M}_{0+} + \epsilon \cdot q \; {\cal H}_{0+} + \ldots \big\} \chi_i \; .\nonumber \end{eqnarray} In these equations all variables have to be expressed in the {\it cm} frame, in particular the components of the 4-momentum transfer, $q = (\omega, \vec{q})$, the polarization vector of the virtual photon, $\epsilon = (\epsilon_0, \vec{\epsilon})$ and the nucleon spin $\vec{\sigma}$. Furthermore, $Q^2 = - q^2$ and $\vec{\epsilon}_T$ is the polarization vector transverse to the direction of the virtual photon. The ellipses denote $P$-waves and higher multipoles. Since the polarization vector $\epsilon_{\mu}$ is proportional to the transition current of the electron, $j_{\mu}$ or $j_{\mu}^5$, the four-product $\epsilon \cdot q$ vanishes exactly for the (conserved) vector current. However, it can also be safely neglected in the case of the axial vector, because the divergence of the axial current is proportional to the mass of the electron. The threshold values of the $S$-wave multipoles have been predicted by general principles following the arguments of low energy theorems (LET). Since these theorems assume Lorentz and gauge invariance and the PCAC relation, they should be obeyed by our model, too. However, there have been recently reported large modifications to LET due to loop corrections. Concerning the vector current these corrections are particular large for neutral pion photoproduction at threshold \cite{Me91}, but do not play an important role for the inclusive cross section, which is dominated by charged pion production. However, it is interesting that two of the multipoles for the axial vector have contributions containing the $\pi N$ $\sigma$-term. According to \cite{Me94} these are \begin{eqnarray} \label{mul} {\cal L}_{0+}^{(+)} &=& \frac{1}{3\pi m_{\pi} f_{\pi}} \left[ \sigma (q^2 - m_{\pi}^2) - \frac{1}{4} \sigma (0) \right] - \frac{a^+ f_\pi}{m_\pi} + O (m_{\pi})\, , \\ {\cal H}_{0+}^{(+)} &=& \frac{a^+ f_{\pi}}{q^2 - m_{\pi}^2} + \frac{\sigma (0) - \sigma (q^2 - m_{\pi}^2)}{12 \pi f_{\pi} (q^2 - m_{\pi}^2)} + O (m_{\pi})\, , \nonumber \end{eqnarray} where $f_{\pi} = 93$ MeV is the pion decay constant and $a^+$ the isospin even $S$-wave $\pi N$-scattering length. As we see from (\ref{mulzer}) and the above considerations, the multipole ${\cal H}_{0+}$ does not contribute in the limit of a vanishing lepton mass. The longitudinal multipole ${\cal L}_{0+}$, however, contributes and its threshold value is dominated by the $\sigma$-term. Unfortunately, it appears in combination with the vector coupling of the $Z^0$ at the vertex of the electron, i.e. this interesting term is suppressed by a factor $(4 \sin^2 \theta_W -1)$. \subsection{Explicit structure of the tensors} In this subsection the explicit structure of the Lorentz tensors will be given. The electromagnetic lepton tensor has the familiar form \begin{eqnarray} \eta_{\mu\nu}^{EM} &=& \eta_{\mu\nu}^{(s)} + h \eta_{\mu\nu}^{(a)} \; , \end{eqnarray} where $h$ denotes the helicity of the incoming electron. The antisymmetric part can be omitted for unpolarized nucleons, because the electromagnetic hadronic tensor is symmetric in this case. The interference tensor for the lepton is \begin{eqnarray} \eta_{\mu\nu}^{I} &=& g_{\mbox{\tiny{\sl V}}}^e (\eta_{\mu\nu}^{(s)} +h \eta_{\mu\nu}^{(a)} ) + g_{\mbox{\tiny{\sl A}}}^e (\eta_{\mu\nu}^{(a)} + h \eta_{\mu\nu}^{(s)}) \\ &=& \eta_{\mu\nu}^{I,(s)} + \eta_{\mu\nu}^{I,(a)} , \nonumber \end{eqnarray} where $g_{\mbox{\tiny{\sl V}}}^e$ and $g_{\mbox{\tiny{\sl A}}}^e$ denote the weak neutral current couplings of the electron \begin{eqnarray} g_{\mbox{\tiny{\sl V}}}^e &=& \frac{1}{4 \sin \theta_W \cos \theta_W} (-1 + 4 \sin^2 \theta_W) \, , \\ g_{\mbox{\tiny{\sl A}}}^e &=& \frac{1}{4 \sin \theta_W \cos \theta_W} \, . \nonumber \end{eqnarray} Using current conservation, the electromagnetic tensor of the nucleon has the Lorentz structure \begin{eqnarray} W_{\mu\nu}^{EM} &=& -g_{\mu\nu} W_1^{EM} + P^i_\mu P^i_\nu \frac{W_2^{EM}}{m^2} + k^\pi_\mu k^\pi_\nu \frac{W_3^{EM}}{m^2} - \frac{1}{2} ( P^i_\mu k^\pi_\nu + P^i_\nu k^\pi_\mu ) \frac{1}{k_\pi \cdot q \; P_i \cdot q \; m^2} \\ & & (q^2 m^2 W_1^{EM} + ( P_i \cdot q)^2 W_2^{EM} \vphantom{\frac{1}{2}} + ( k_\pi \cdot q )^2 W_3^{EM} - q^4 W_4^{EM} ) \; . \vphantom{\frac{1}{2}} \nonumber \end{eqnarray} The symmetric part of the hadronic interference tensor $W_{\mu\nu}^{I}$ has the same structure as $W_{\mu\nu}^{EM}$. The corresponding antisymmetric part is \begin{eqnarray} \label{Wia} W_{\mu\nu}^{I,(a)} &=& -i \epsilon_{\mu\nu\alpha\beta} P_i^\alpha q^\beta \frac{W_5^{I}}{m^2} -i \epsilon_{\mu\nu\alpha\beta} k_\pi^\alpha P_i^\beta \frac{W_{6}^{I}}{m^2} -i \epsilon_{\mu\nu\alpha\beta} k_\pi^\alpha q^\beta \frac{W_{7}^{I}}{m^2} -i ( P^i_\mu k^\pi_\nu - P^i_\nu k^\pi_\mu ) \frac{W_{8}^I}{m^2} \\ & & -i ( P^i_\mu \epsilon_{\nu\alpha\beta\gamma} - P^i_\nu \epsilon_{\mu\alpha\beta\gamma} ) k_\pi^\alpha q^\beta P_i^\gamma \frac{W_{9}^I}{m^4} -i ( k^\pi_\mu \epsilon_{\nu\alpha\beta\gamma} - k^\pi_\nu \epsilon_{\mu\alpha\beta\gamma} ) k_\pi^\alpha q^\beta P_i^\gamma \frac{W_{10}^I}{m^4} \, . \nonumber \end{eqnarray} As the axial currents are not conserved, there are in principle additional terms proportional to $q_\mu$ or $q_\nu$. Since these terms vanish after contraction with the leptonic tensor (\ref{lept}), they have been omitted in (\ref{Wia}) right away. The structure functions $W_j(\nu,\nu_B,q^2)$ can be expressed in terms of the invariant amplitudes (\ref{ia}) in a straightforward way. The result of this calculation is given in Appendix \ref{strf}. \section{Model for the Hadronic Currents} \label{mod} In this section we will present the phenomenological model that is used to calculate the hadronic currents. It contains contributions of the Born terms and a phenomenological description of the $\Delta(1232)$ resonance. \subsection{Nonresonant contributions} The background of the Born terms contains both vector current and axial vector current contributions. Since we neglect the strangeness of the nucleon, the weak vector current differs from the electromagnetic current only by a coupling constant. Accordingly, we have to calculate the Feynman diagrams of Fig. \ref{fig3}. While all diagrams contribute to the vector current, the pion pole diagram does not contribute to the axial current. For the calculation of the Feynman diagrams we use the following interaction Lagrangians \begin{eqnarray} \label{L} {\cal L}_{\pi N N}^{PV} &=& \frac{f_{\pi N N}}{m_\pi} \bar{\psi} \gamma_\mu \gamma_5 \vec{\tau} \psi \cdot \partial^\mu \vec{\phi} \, , \\ {\cal L}_{V^\mu N N} &=& - e \, \bar{\psi} \frac{1}{2} \bigg[ (\xi_V^{I=0} F_1^s + \xi_V^{I=1} \tau_3 F_1^v) \gamma_\mu V^\mu - (\xi_V^{I=0} \kappa_s F_2^s + \xi_V^{I=1} \kappa_v \tau_3 F_2^v) \frac{\sigma_{\mu\nu}}{2 m} \partial^\nu V^\mu \bigg] \psi \, , \nonumber \\ {\cal L}_{V^\mu \pi \pi} &=& -e\,\xi_V^{I=1} F_\pi ( \vec{\phi} \times \partial_\mu \vec{\phi} )_3 V^\mu \, , \nonumber \\ {\cal L}_{V^\mu N N \pi} &=& e \, \xi_V^{I=1} \frac{f_{\pi N N}}{g_a m_\pi} G_A \bar{\psi} ( \vec{\tau} \times \vec{\phi} )_3 \gamma_\mu \gamma_5 \psi \, V^\mu , \, \nonumber \\ {\cal L}_{A^\mu N N} &=& - e \, \xi_A^{I=1} \bar{\psi} \left[ G_A \gamma_\mu A^\mu + G_P \frac{i \partial_\mu}{2 m} A^\mu \right] \gamma_5 \frac{\tau_3}{2} \psi \, , \nonumber \\ {\cal L}_{A^\mu N N \pi} &=& e \, \xi_A^{I=1} \frac{f_{\pi N N}}{ g_a m_\pi} \bar{\psi} ( \vec{\tau} \times \vec{\phi} )_3 \left( F_1^v \gamma_\mu A^\mu - \kappa_v F_2^v \frac{ \sigma_{\mu\nu}}{2 m} \partial^\nu A^\mu \right) \psi \, , \nonumber \end{eqnarray} with $\kappa_{(s,v)} = (\kappa_p \pm \kappa_n)$. The isospin factors $\xi$ follow from the decomposition of the corresponding quark current operators according to strong isospin \cite{Mu94}. For the electromagnetic current these factors are equal to unity, for the weak neutral current we have \begin{eqnarray} \xi_V^{I=1} &=& \frac{1}{2 \sin \theta_W \cos \theta_W} ( 1 - 2 \sin^2 \theta_W ) \, ,\\ \xi_V^{I=0} &=& -\frac{1}{2 \sin \theta_W \cos \theta_W} 2 \sin^2 \theta_W \, , \nonumber \\ \xi_A^{I=1} &=& -\frac{1}{2 \sin \theta_W \cos \theta_W} \, . \nonumber \end{eqnarray} Note that there appears no isoscalar contribution to the axial vector current, i.e. $\xi_A^{I=0} = 0$, because strange quarks have been neglected. The form factors $F_1$, $F_2$, $F_\pi$, $G_A$ and $G_P$ are functions of momentum transfer $q^2$. Since our calculation will be performed at relatively small momentum transfer, we have used simple dipole forms for the Sachs form factors, with the assumption $G_A/g_a = F_\pi = F_1^v$. This insures gauge invariance without additional gauge terms. With standard methods and neglecting the lepton mass terms, we obtain the invariant matrix elements \begin{eqnarray} \label{mat_b} {\cal M}_s &=& \xi_V^{I=1} \frac{f_{\pi N N}}{m_\pi} \bar{u}(\vec{P_f}) \bigg[ \gamma_5 \dida{k_\pi} \frac{\didag{P_i} + \dida{q} + m}{s - m^2} \left( \tilde I_s^D \dida{\epsilon} + i \tilde I_s^P \frac{\sigma_{\mu\nu}}{2 m} q^\nu \epsilon^\mu \right) \bigg] u(\vec{P_i})\, , \\ {\cal M}_u &=& \xi_V^{I=1} \frac{f_{\pi N N}}{m_\pi} \bar{u}(\vec{P_f}) \bigg[ \left( \tilde I_u^D \dida{\epsilon} + i \tilde I_u^P \frac{\sigma_{\mu\nu}}{2 m} q^\nu \epsilon^\mu \right) \frac{\didag{P_i} - \dida{k_\pi} + m}{u - m^2} \gamma_5 \dida{k_\pi} \bigg] u(\vec{P_i})\, , \nonumber \\ {\cal M}_t &=& \xi_V^{I=1} \frac{f_{\pi N N}}{m_\pi} \frac{\epsilon \cdot ( 2 k_\pi - q)}{t - m_\pi^2} 2 m I_t F_\pi \bar{u}(\vec{P_f}) \gamma_5 u(\vec{P_i}) \, , \nonumber \\ {\cal M}_{c} &=& \xi_V^{I=1} \frac{f_{\pi N N}}{g_a m_\pi} I_t G_A \bar{u}(\vec{P_f}) \dida{\epsilon} \gamma_5 u(\vec{P_i}) \, , \nonumber \\ {\cal M}_s^{5} &=& -\xi_A^{I=1} \frac{f_{\pi N N}}{m_\pi} \bar{u}(\vec{P_f}) \left[ \dida{k_\pi} \frac{\didag{P_i} + \dida{q} - m}{s - m^2} \tilde I_s^A \dida{\epsilon} \right] u(\vec{P_i})\, , \nonumber \\ {\cal M}_u^{5} &=& -\xi_A^{I=1} \frac{f_{\pi N N}}{m_\pi} \bar{u}(\vec{P_f}) \left[ \tilde I_u^A \dida{\epsilon} \frac{\didag{P_i} - \dida{k_\pi} - m}{u - m^2} \dida{k_\pi} \right] u(\vec{P_i})\, , \nonumber \\ {\cal M}_{c}^{5} &=& \xi_A^{I=1} \frac{f_{\pi N N}}{g_a m_\pi} I_t \bar{u}(\vec{P_f}) \left[ \gamma_\mu F_1^v + i \frac{F_2^v}{2 m} \sigma_{\mu\nu} q^\nu \right] \epsilon^\mu u(\vec{P_i})\, . \nonumber \end{eqnarray} Note that there do not appear any induced pseudoscalar terms, because the lepton mass terms have been neglected. The electromagnetic matrix element is the sum of the first 4 terms in (\ref{mat_b}), corresponding to s, u and t channel pole terms and the contact term(''c''). The matrix element of the weak neutral current is the corresponding sum of the combinations ${\cal M} +{\cal M}^5$. The following abbreviations for the isospin matrix elements have been used in (\ref{mat_b}): \begin{eqnarray} I_t &=& I^- , \\ \tilde I_{(s,u)}^D &=& \frac{1}{2} \left[ \frac{\xi_V^{I=0}}{\xi_V^{I=1}} F_1^s I^0 + F_1^v ( I^+ \pm I^-) \right], \nonumber \\ \tilde I_{(s,u)}^P &=& \frac{1}{2} \left[ \frac{\xi_V^{I=0}}{\xi_V^{I=1}} F_2^s I^0 + F_2^v ( I^+ \pm I^-) \right], \nonumber \\ \tilde I_{(s,u)}^A &=& \frac{1}{2} \left[ G_A ( I^+ \pm I^-) \right] \, , \nonumber \end{eqnarray} with $I^{(\pm,0)}$ as defined in (\ref{iso}). The invariant amplitudes obtained from (\ref{mat_b}) are identical to the results of \cite{Ad68}. They can be found in Appendix \ref{iam_b}. \subsection{Resonant contributions} In the kinematical region between threshold and about 400 MeV excitation energy, the dominant resonant contributions are due to the $\Delta(1232)$. We treat the $\Delta$ as a Rarita-Schwinger field and use the on-shell form of the propagator modified by a phenomenological constant width, \begin{eqnarray} \label{d_prop} \Delta_{\mu\nu} &=& \frac{\dida{p} + M}{p^2 -M^2 + i \Gamma M } \left( g_{\mu\nu} -\frac{1}{3} \gamma_\mu \gamma_\nu -\frac{2 p_\mu p_\nu}{3 M^2} +\frac{p_\mu \gamma_\nu -\gamma_\mu p_\nu}{3 M} \right) \, . \end{eqnarray} $M = 1210 \mbox{ MeV}$ is obtained as the real part of the $\Delta$ pole \cite{PDG94}. The width $\Gamma = 85 \mbox{ MeV}$ is fitted to the experimental data for pion electroproduction \cite{Co67,Ly67}, which is close to the imaginary part of the pole \cite{PDG94}, $\Gamma = 100 \mbox{ MeV}$ . The Feynman diagrams to be calculated are shown in Fig. \ref{fig4}. We use the phenomenological $N \Delta$ transition currents \cite{Na82} \begin{eqnarray} {J}_{\mu}^{EM} &=& \bar{u}^{\lambda}(\vec{p'}) \left[ \left( \frac{C_3^{\gamma}}{m} \gamma^{\nu} + \frac{C_4^{\gamma}}{m^2} {p'}^{\nu} + \frac{C_5^{\gamma}}{m^2} p^{\nu} \right) (g_{\lambda\mu} g_{\rho\nu} - g_{\lambda\rho} g_{\mu\nu}) q^{\rho} \gamma_5 \right] u(\vec{p})\, , \\ {J}_{\mu}^{NC} &=& \bar{u}^{\lambda}(\vec{p'}) \left[ \left( \frac{C_{3V}^{Z}}{m} \gamma^{\nu} + \frac{C_{4V}^{Z}}{m^2} {p'}^{\nu} + \frac{C_{5V}^{Z}}{m^2} p^{\nu} \right) \vphantom{\frac{1}{2}} (g_{\lambda\mu} g_{\rho\nu} - g_{\lambda\rho} g_{\mu\nu}) q^{\rho} \gamma_5 + C_{6V}^Z g_{\lambda\mu} \gamma_5 \right. \nonumber \\ & &+ \left. \left( \frac{C_{3A}^{Z}}{m} \gamma^{\nu} + \frac{C_{4A}^{Z}}{m^2} {p'}^{\nu} \right) (g_{\lambda\mu} g_{\rho\nu} - g_{\lambda\rho} g_{\mu\nu}) q^{\rho} + C_{5A}^Z g_{\lambda\mu} + \frac{C_{6A}^{Z}}{m^2} p_{\lambda} q_{\mu} \right] u(\vec{p}) \, . \nonumber \end{eqnarray} The form factors $C_3$ to $C_6$ are taken from Nath et al.\cite{Na82}, with a $q^2$ dependence as in \cite{La88}. The $\pi N \Delta$ vertex has the effective form \cite{Ti78} \begin{equation} \label{piND} \Lambda^{\mu}_{\pi N \Delta} = \frac{f_{\pi N \Delta}}{m_{\pi}} k_\pi^{\mu}\, . \end{equation} With (\ref{d_prop})-(\ref{piND}) we obtain the invariant matrix elements for s channel $\Delta$ production, \begin{eqnarray} {\cal M}^{EM}_{s \Delta} &=& \frac{f_{\pi N \Delta}}{m m_{\pi}} \bar{u}(\vec{P_f}) k^{\pi}_{\sigma} \Delta^{\sigma \lambda} \bigg[ C_3^{\gamma}( q_\lambda \gamma_\mu - g_{\lambda \mu} \dida{q}) +\frac{C_4^{\gamma}}{m}( q_\lambda(P_f + k_\pi)_{\mu} \\ & & -q \cdot ( P_f + k_\pi) g_{\lambda \mu} ) \bigg] \gamma_5 \epsilon^\mu u(\vec{P_i}) \, , \nonumber \\ {\cal M}^{NC}_{s \Delta} &=& \xi_V^{I=1} {\cal M}^{EM}_{s \Delta} \vphantom{\frac{1}{2}}, \nonumber \\ {\cal M}^{5 NC}_{s \Delta} &=& -\xi_A^{I=1} \frac{f_{\pi N \Delta}}{m m_{\pi}} \bar{u}(\vec{P_f}) k^{\pi}_{\sigma} \Delta^{\sigma \lambda} \bigg[ C_3^{A}( q_\lambda \gamma_\mu - g_{\lambda \mu} \dida{q}) +\frac{C_4^{A}}{m}( q_\lambda(P_f + k_\pi)_{\mu} \nonumber \\ & & -q \cdot ( P_f + k_\pi) g_{\lambda \mu} ) - C_5^{A} m g_{\lambda \mu} \bigg] \epsilon^\mu u(\vec{P_i}) \, . \nonumber \end{eqnarray} The corresponding u channel amplitudes are constructed by using the crossing properties of the invariant amplitudes (\ref{cross}). These invariant matrix elements can now be decomposed into invariant amplitudes as in (\ref{ia}). The result of this calculation is listed in Appendix \ref{iam}. \section{Results} \label{res} As a first test of the model we have calculated the inclusive electromagnetic cross section. The experimental data for this process \cite{Co67,Ly67} could be reproduced within about 5\% over the whole energy region from threshold to about 400 MeV excitation energy. Fig. \ref{fig5} shows the asymmetry for the proton and the neutron in various kinematical situations. The two upper figures compare the asymmetry for proton and neutron as function of the photon equivalent energy, \begin{equation} k_{\gamma} = (W^2 - m^2)/2m \, , \end{equation} the {\it lab} energy necessary to excite a hadronic system with {\it cm} energy W. At small excitation energies the asymmetry is essentially given by background contributions, whereas the $\Delta$ resonance dominates at the higher energies. While the individual contributions vary strongly, the coherent sum is essentially constant. In particular, the resonance structure of the inclusive cross sections is essentially wiped out when the asymmetries are calculated. As demonstrated in the two lower figures of Fig. \ref{fig5}, this effect is quite independent of the kinematics except for a trivial dependence on the momentum transfer. For the proton this $Q^2$ dependence is well described by the simple estimate \begin{equation} \label{gr_a} A \approx -\frac{G_F Q^2}{e^2} \approx -10^{-4} Q^2/\mbox{ GeV}^2 \; , \end{equation} where $G_F$ is the Fermi coupling constant. In the case of the neutron, the asymmetry is somewhat smaller, at about 80\% of the proton case, showing a slight enhancement in the resonance peak. In Fig. \ref{fig6} the asymmetry is shown at an incident electron energy of $\epsilon_i = 800 \mbox{ MeV}$ and at $W = 1232 \mbox{ MeV}$, directly in the resonance peak. In this kinematical situation, the asymmetry is strongly dominated by the $\Delta$, the background contributions cancel against the interference terms (upper panel). Therefore, the neglect of background terms as in the work of Nath et al.\cite{Na82} is well justified. As is shown in the lower panel of this figure, we are able to reproduce those results except for an overall scaling factor of about 90 \%. In the work of Cahn et al.\cite{Ca78} the vector coupling between the $Z^0$ and the electron has been neglected. This corresponds to $\sin^2 \theta_W = 1/4$ and produces a somewhat lower curve with a flat distribution in $Q^2$. We have also compared our calculation to the results of Li et al. \cite{Li82} and find a reasonable agreement. However, the calculation of Ishankuliev et al. \cite{Is80} disagrees with all others by a factor $1/2$ seemingly due to a wrong coupling constant. Preliminary studies of the $\pi N$ $\sigma$-term (\ref{mul}) have shown only small effects of this physically interesting quantity on the asymmetry. Unfortunately, the coupling of the $Z^0$ to the axial current of the hadron appears together with the vector coupling to the electron which is strongly suppressed by the value of the Weinberg angle. In addition, the $S$-wave multipole ${\cal L}_{0+}$ contributes to the asymmetry only by interference with the electromagnetic $P$-waves. In the usual conventions this contribution is proportional to $(2 M_{1+} + M_{1-}) {\cal L}_{0+}$. As a consequence, the multipole ${\cal L}_{0+}$ is further suppressed near threshold. \section{Summary and Conclusions} \label{conc} The aim of this work has been to investigate parity violating (PV) contributions to pion electroproduction. Only PV effects due to interference between $\gamma$ and $Z^0$ exchange have been considered. PV effects in the strong interaction, as discussed by Li et al.\cite{Li82}, have been neglected. The appropriate observable to study these effects is the asymmetry defined in (\ref{def_a}). A simple phenomenological model with effective Lagrangian densities has been constructed that allows for the calculation of the asymmetry in the kinematical region from pion threshold to $\Delta(1232)$ resonance. The model is fully relativistic, fulfils the crossing symmetry and is gauge invariant due to our simple choice of the form factors. The invariant amplitudes for the nonresonant contributions respect PCAC. The $\Delta$ resonance is treated as a Rarita-Schwinger field with phenomenological transition currents. For the propagator, the on-shell form with the $\Delta$ pole of the $\pi N$ scattering matrix is used. The nonresonant contributions are created by the usual pseudovector $\pi N$ coupling and phenomenological transition currents. The weak neutral vector current is traced back to the electromagnetic one by decomposing the quark currents according to strong isospin. For the axial current, a contact term has been introduced in addition to the $s$ and $u$ channel nucleon Born terms. This contact term is necessary to fulfil the low energy theorem of Adler \cite{Ad68} for the nonresonant contributions. The calculated asymmetry is nearly constant as function of excitation energy. It increases linearly with the square of the four-momentum transfer, $A \approx 10^{-4} q^2 /\mbox{GeV}^2$ for the proton, and has about 80\% of that value for the neutron. Previous calculations for the excitation of the $\Delta$ isobar could be reproduced reasonably well, while our results are at variance with some of the earlier work on the background contributions. In conclusion we find the following results: \begin{itemize} \item The asymmetry grows linearly with the momentum transfer $Q^2$ and is nearly independent of the excitation energy. \item The expected asymmetries for pion electroproduction are comparable with the asymmetries found for elastic electron scattering. \item Because of the value of the Weinberg angle, the vector coupling of the $Z^0$ at the electron vertex is suppressed. As a consequence, the asymmetry is dominated by the hadronic vector current, the contributions of the hadronic axial currents being of the order of 10 - 20 \% only. \item A precision measurement of the asymmetries is potentially an independent experiment to determine the important $\pi N$ $\sigma$-term, which appears in the $S$-wave multipole ${\cal L}_{0+}$ of the hadronic axial current. Unfortunately, the $\sigma$-term will be difficult to measure, because the hadronic axial current is suppressed and furthermore, the multipole ${\cal L}_{0+}$ appears in the asymmetry only by its interference with $P$-wave multipoles. \end{itemize} Finally, we would like to comment on some possible improvements of the calculations: \begin{itemize} \item The simple superposition of Born and resonance terms violates the Watson theorem. It would be necessary to perform a multipole decomposition and to unitarize at least the multipoles carrying the phase of the $\Delta(3,3)$ resonance. \item We have neglected the contribution of the strange sea to the hadronic transition currents. Taking account of such effects would require additional, as yet undetermined form factors. However, it is unlikely that pion production would be strongly modified by strangeness degrees of freedom. \item Though we have not found large contributions of the $\sigma$-term to the asymmetry, this matter deserves more systematical studies. Clearly, any further independent measurement of this important quantity would be invaluable. \end{itemize}
\section{Introduction} The soft particle production in a very high energy hadron-hadron or nucleus-nucleus collision is an interesting phenomenon. Occasionally, the collision creates a large number of low energy (small $p_t$) particles, mainly pion quanta, initially populating in a small interaction volume and subsequently undergoing a rapid expansion. The perturbative QCD is not applicable in describing the dynamics since it involves a large number of quanta and the interactions are highly nonlinear. One may anticipate some novel dynamical feature of the nonperturbative QCD. Although there is some evidence that the $p_t$ distribution of these particles follows the scaling law with an effective temperature, it is not clear whether or not these low energy particles can actually thermalize so that their distribution can be described by thermodynamics. In fact, some deviation from the thermal distribution in the very small $p_t$ region, say, $p_t<100$ MeV, has been observed though data are poor in this region at present. On the other hand, it has been suggested, first by Horn and Silver \cite{horn}, that these low energy pions may be described by a classical theory. The number ($N$) of quanta involved is large, the quantum fluctuation is suppressed by $1/\sqrt{N}$. In addition, the low energy theorem on the pion-pion scatterings dictates that the quantum corrections to the scattering amplitudes are suppressed by a factor of $p^2/(4\pi f_\pi)^2$. More recently, a scenario of disoriented chiral condensate \cite{DCC,rw} suggests that these low energy pions may be out of equilibrium and undergo a quench following a chiral phase transition, and their interactions should be described by the classical chiral dynamics. In this paper, we shall determine the possible classical field evolutions that these low energy pions may follow based on the nonlinear $\sigma$-model. The advantage of the {\it nonlinear} $\sigma$-model over the {\it linear} $\sigma$ model is that the constraint of the vacuum expectation value on the fields is built in and that the pion fields always describe the massless modes irrespective of the vacuum orientation in the background. The $\sigma$ mass is taken to be infinity so the low energy structure of the theory is evident. We have obtained in an analytic form a class of classical solutions to the nonlinear $\sigma$-model in $3+1$ space-time dimensions as the candidates of the disoriented chiral condensate in QCD. Our general solutions have a transverse momentum distribution and need not be subject to a boost-invariance constraint. The solution with a nonuniform isospin orientation is constructed by the chiral $SU(2)\times SU(2)$ rotation from a uniformly oriented solution. In the limit of a boost invariance and no transverse momentum, our solutions reduce to those of Blaizot and Krzywicki \cite{blaiz}. We study the distribution of the neutral pion fraction $f$ for the pions that disintegrate from the disoriented vacua. We find that the distribution dP/df = 1/(2$\sqrt{f})$ holds for the uniformly oriented vacua and also for the boost-invariant vacua with an infinitely large uniform spread in the transverse direction, but it does not hold for the vacuum whose isospin orientation is nonuniform in space-time. However this distribution should be correct if one selects pions from within a small region in the $y$-${\bf k}_\bot$ plot event by event. This conclusion is reached through the analysis using the classical field theory method and also by studying the quantum pion states. We organize the paper as follows. In Sec.\ 2, we start with the analysis for the boost-invariant solutions with no transverse momentum. In Sec.\ 3, we make the observation that all solutions with a nonuniform isospin orientation are obtained by the chiral rotations from a uniformly oriented solution whose energy is degenerate with the nonuniform ones. In Sec.\ 4, we give a general solution with a uniform isospin orientation, from which we can obtain the nonuniform solutions by the chiral rotations according to the prescription given in Sec.\ 3. The general solution has a nontrivial transverse momentum distribution and is not subject to the boost-invariance constraint. In Sec.\ 5, we examine the charge distribution of the pions disintegrating from these disoriented vacua. The picture of classical field theory leading to the distribution dP/df = 1/(2$\sqrt{f})$ does not apply to the nonuniformly oriented vacua except in the boost invariant limit with zero transverse momentum. For a general solution which has rapidity and transverse momentum dependence, the distribution holds only within each small segment in the $y$-${\bf k}_\bot$ plot. The charge distribution is also studied from the viewpoint of quantum multipion states, following Horn and Silver \cite{horn}. The modification of the distribution is attributed to the fact that for the general nonuniform solution, more than one orbital state is available for pions to occupy so that there are many different ways to construct multipion states. \section{Boost-Invariant Solution in $1+1$ Dimensions} In high energy hadron or nucleus collisions, the configurations approximately invariant along the collision axis are of particular interest. We first focus on this class of solutions ignoring the transverse spatial dependence. We choose a {\it nonlinear} $\sigma$-model as the dynamical model for QCD at low energy. The solutions that we obtain in this Section are equivalent to those of Blaizot and Krzywicki $\cite{blaiz}$ though they are dervied in a slightly different way in order to clarify a relation between the uniformly oriented solutions and the nonuniformly oriented ones which plays an important role when we extend our argument to the more general case later. The phase and radial representation of the nonlinear $\sigma$-model is, \begin{equation} \Sigma (x) = e^{i{\mbox{\boldmath $\tau$} \cdot{\bf n}(x)\theta(x)}}. \end{equation} No matter what values the classical phase fields take, the state remains at the bottom of the potential valley because $|\Sigma| = 1$. This facilitates greatly the search for the DCC-type solutions which are realized at the bottom of potential well. Define the pion field \begin{equation} \mbox{\boldmath $\pi$} (x) = f_{\pi}{\bf n}(x)\,\theta(x). \end{equation} where ${\bf n}(x)$ is an unit isovector field obeying ${\bf n}(x)\cdot{\bf n} (x) = 1$. Alternatively one defines $\mbox{\boldmath $\pi$} (x)$ by $\sigma + i\mbox{\boldmath $\tau\cdot\pi$} = f_{\pi}\Sigma$ with the constraint $\sigma =\sqrt{f_{\pi}^2 - \mbox{\boldmath $\pi$}^2}$. In this case the pion fields are given by $\mbox{\boldmath $\pi$}(x) = f_{\pi}{\bf n}(x)\:{\rm sin}\theta(x)$. In either case, ${\bf n}$ determines the isospin orientation of the pion field. The lagrangian is given by \begin{equation} {\cal L} = \frac{f_{\pi}^2}{4} {\rm tr} \Bigl(\partial_{\mu}\Sigma^\dagger(x) \partial^{\mu}\Sigma(x)\Bigr), \nonumber \end{equation} where $\Sigma$ transforms like $\Sigma\rightarrow U_L\Sigma U_R^\dagger$ under $SU(2)_L\times SU(2)_R$ rotations. In terms of $\theta(x)$ and ${\bf n}(x)$, the lagrangian is \begin{equation} {\cal L} = \frac{f_{\pi}^2}{2}(\partial_{\mu}\theta\:\partial^{\mu}\theta + \sin^{2}\theta\partial_{\mu}{\bf n}\cdot\partial^{\mu}{\bf n}) + \frac{\lambda f_{\pi}^2}{2}({\bf n}^2 -1), \end{equation} where $\lambda$ is a Lagrange multiplier. We will not include an explicit chiral symmetry breaking throughout this paper. The Euler-Lagrange equations are \begin{equation} \Box\theta = \sin\theta\: \cos\theta\:\partial_{\mu}{\bf n}\cdot\partial^{\mu} {\bf n}, \end{equation} \begin{equation} \partial_{\mu}(\sin^{2}\theta\:\partial^{\mu}{\bf n}) = \lambda{\bf n}. \end{equation} The chiral $SU(2)_{L}\times SU(2)_{R}$ symmetry assures the conservation of the vector and axial-vector currents. In terms of $\theta$ and ${\bf n}$, the current conservation is written as \begin{equation} \partial_{\mu}(\sin^{2}\theta\:{\bf n}\times\partial^{\mu}{\bf n}) = 0, \end{equation} \begin{equation} \partial_{\mu}({\bf n}\:\partial^{\mu}\theta + \sin\theta\: \cos\theta\:\partial^{\mu}{\bf n}) = 0. \end{equation} The isospin current conservation (7) follows also from (6), while the axial-vector current conservation (8) can be derived from (5) by repeated use of $({\bf n}\cdot d{\bf n}/d\tau) = 0$. (4), (5), (6) and (7) are most general with no assumptions or approximations made. We consider a boost-invariant case in 1+1 dimensions where the fields $\theta(x)$ and ${\bf n}(x)$ are only functions of the variable $\tau$: \begin{equation} \tau = \sqrt{t^{2}-x^{2}}. \end{equation} For a function only of $\tau$, a partial derivative $\partial_{\mu}f(\tau)$ is equal to $(x_{\mu}/\tau)df/d\tau$. Furthermore, $\partial_{\mu}(f(\tau)\partial^{\mu}g(\tau)) = (1/\tau^2)(d(\tau^{2}fg^\prime)/d\tau)$ where $g^\prime = dg/d\tau$. The current conservation relations can be integrated into \begin{equation} \tau \: \sin^{2}\theta\:{\bf n}\times\frac{d{\bf n}}{d\tau}= {\bf a}, \end{equation} \begin{equation} \tau\,{\bf n}\frac{d\theta}{d\tau}+\tau\,\sin\theta\:\cos\theta\: \frac{d{\bf n}}{d\tau}= {\bf b}, \end{equation} where ${\bf a}$ and ${\bf b}$ are constant vectors in the isospin space whose magnitudes are denoted as $a$ and $b$ respectively. It is immediately obvious from (10) and (11) that ${\bf a}$ and ${\bf b}$ are orthogonal to each other: \begin{equation} {\bf a}\;\bot\;{\bf b}. \end{equation} The isovector field ${\bf n}(\tau)$ stays perpendicular to ${\bf a}$ as $\tau$ varies. By multiplying (10) with ${\bf n}$ vectorially and using ${\bf n}\cdot (d{\bf n}/d\tau) = 0$, one obtains \begin{equation} \frac{d{\bf n}}{d\tau} = \frac{{\bf a}\times{\bf n}}{\tau\sin^{2}\theta}, \end{equation} a standard equation for a vector ${\bf n}$ to precess around a constant vector ${\bf a}$. The precession frequency $|{\bf a}|/\tau\sin^2\theta$ varies with the proper time $\tau$. The relations among ${\bf a}$, ${\bf b}$ and ${\bf n}$ are illustrated in Figure 1. Squaring (11) gives \begin{equation} \Biggl(\tau\frac{d\theta}{d\tau}\Biggr)^{2}+ \sin^{2}\theta\:\cos^{2}\theta\ \Biggl(\tau\frac{d{\bf n}}{d\tau}\Biggr)^{2} = b^{2}. \end{equation} Eliminating $d{\bf n}/d\tau$ from the these equations, one obtains the differential equation for $\theta(\tau)$: \begin{equation} \Biggl(\tau\frac{d\theta}{d\tau}\Biggr)^{2} = a^{2}+ b^{2} - \frac{a^{2}}{\sin^{2}\theta}, \end{equation} where $a = |{\bf a}|$ and $b = |{\bf b}|$. (13) and (15) combined contain the same information as the first integrals of the Euler-Lagrange equations for $\theta$ and ${\bf n}$ so that we may proceed with the current conservation laws. (15) is analytically integrable into the most general boost-invariant solution for $\theta(\tau)$ in 1+1 dimension: \begin{equation} \cos\theta(\tau) = (b/\kappa)\; \cos\Bigl(\kappa\:{\rm ln}(\tau/\tau_0) + \vartheta_0\Bigr), \end{equation} where $\kappa^2 = a^2+b^2$ and $ \cos\vartheta_0 =(\kappa/b)\,\cos \theta (\tau_0)$. Substituting $\theta(\tau)$ in the integral from the isovector current conservation, one obtains the general solution for ${\bf n}(\tau)$. Since ${\bf n}(\tau)$ and ${\bf b}$ both lie in the plane perpendicular to ${\bf a}$, it is convenient to express the unit isovector field ${\bf n}(\tau)$ in terms of a single angle $\beta(\tau)$ measured from the direction of ${\bf b}$: \begin{equation} {\bf n}(x)\cdot{\bf b} = b \cos\beta(x). \end{equation} The solution for $\cos\,\beta(\tau)$ can be expressed in terms of $\theta(\tau)$ \begin{equation} \cos\,\beta(\tau) =\frac{a}{b}\sqrt{\frac{\kappa^2}{a^2} -\frac{1}{\sin^{2}\theta(\tau)}}. \end{equation} The DCC configuration may be obtained from the boost-invariant solution multiplied by a step function ${\rm\Theta}(\tau^2)$: \begin{equation} \Sigma(x)_{DCC} = e^{i\theta (x)\mbox{\boldmath $\tau$ } \cdot{\bf n}(x)}{\Theta}(\tau^2), \end{equation} such that the causality condition is satisfied. Our solutions should only apply to the inside of the light-cone. Once $\Theta(\tau^2)$ is inserted, there appears a source on the light cone at $\tau = 0$ which triggers the formation of a DCC. The energy density ${\cal E}(x)$ of our solution is singular as we approach the light cone: \begin{equation} {\cal E}(x) = \frac{f_{\pi}^2}{2} \Biggl(\frac{t^2+z^2}{\tau^2}\Biggr)(a^2+b^2), \end{equation} for both the uniform and nonuniform solutions. The isospin vectors ${\bf a}$ and ${\bf b}$ enter the energy density in the combination of $a^2+b^2$ as required by $SU(2)\times SU(2)$ invariance. The lowest energy solution of $a^2+b^2 =0$ is a trivial solution obtained from (16), (17) and (18) by taking the limit of $a, b \rightarrow 0$: \begin{equation} \theta(x) = {\rm constant},\; {\bf n}(x) = {\rm constant\;vector}. \end{equation} Blaizot and Krzywicki \cite{blaiz} expressed the pion fields by $\pi = f_{\pi}{\bf n}\,\sin\theta$. If we choose our $\it{initial}$ condition such that $\sin\,\vartheta_0 = 0$, that is, $\cos\theta(\tau_0)=b/\kappa$, our general solution given by (16) and (18) coincides with theirs. \section{Chiral Rotation and Nonuniform Solution} In this Section we focus on the relation between the space-time dependence of the isovector field ${\bf n}$(x) and $SU(2)\times SU(2)$ invariance of the lagrangian. It is easy to see in (18) that our solutions have a uniform isospin orientation when the isospin vector ${\bf a}$ vanishes. When ${\bf a} =0$, $\beta(\tau) = 0\;({\rm mod}\; 2\pi)$, that is, ${\bf n}$ points to the direction of ${\bf b}$ for all uniform DCC's. Because of the $SU(2)\times SU(2)$ invariance, it is always possible to rotate the vector ${\bf a}$ by an appropriate axial rotation to the direction of the vector ${\bf b}$. After the rotation the solution has a uniform isospin orientation and is degenerate in energy with the nonuniform solution prior to the rotation. It is not unfamiliar that if a system possesses some symmetry, a set of infinitely many new solutions may be obtained by making the symmetry transformation on a single solution. To be explicit in the present case, let us rotate a uniform solution \begin{equation} \Sigma_0 (x) = e^{i\theta_0(x)\mbox{\boldmath $\tau$}\cdot{\bf n}_0}, \end{equation} where ${\bf n}_0$ is space-time independent. Upon a general chiral rotation parametrized by $U_L = e^{i\xi\mbox{\boldmath $\tau$} \cdot{\bf n}_L}$ and $U_R = e^{i\eta\mbox{\boldmath $\tau$} \cdot{\bf n}_R}$, the uniform solution is rotated into $\Sigma (x) = U_L\Sigma_0 (x) U_R^{-1}$. The transformed $\theta$ and ${\bf n}$ fields are given by \begin{eqnarray} c_\theta & = & \Bigl(c_\xi c_\eta + ({\bf n}_L\cdot{\bf n}_R)s_\xi s_\eta \Bigr) c_0 + \nonumber \\ & & \Bigl(({\bf n}_0\cdot{\bf n}_R)c_\xi s_\eta - ({\bf n}_0\times{\bf n}_L) s_\xi c_\eta +({\bf n}_0\times {\bf n}_L)\cdot{\bf n}_R\, s_\xi s_\eta\Bigr)s_0, \end{eqnarray} and \begin{eqnarray} {\bf n} s_\theta & = &\Bigl({\bf n}_L s_\xi c_\eta -{\bf n}_R c_\xi s_\eta +({\bf n}_L\times{\bf n}_R)s_\xi s_\eta\Bigr) c_0 \nonumber \\ & & \mbox{} + \Biggl({\bf n}_0\, c_\xi c_\eta + ({\bf n}_0\times{\bf n}_L)s_\xi c_\eta + ({\bf n}_0\times{\bf n}_R)c_\xi s_\eta \nonumber \\ & & \mbox{} +\Bigl(({\bf n}_0\cdot{\bf n}_L){\bf n}_R +({\bf n} _0\cdot{\bf n}_R){\bf n}_L -({\bf n}_L\cdot{\bf n}_R){\bf n}_0\Bigr)s_\xi s_\eta\Biggr)s_0, \end{eqnarray} where $c_{\theta}$ and $s_{\theta}$ stand for $\cos\theta$ and $\sin\theta$, respectively, and so forth, while $c_0 = \cos\theta_0$ and $s_0 = \sin \theta_0$. For an isospin rotation, we choose $\xi = \eta$ and ${\bf n}_L = {\bf n}_R$. Then the rotated fields are \begin{eqnarray} \cos\theta(\tau)&=& \cos\theta_0(\tau), \nonumber \\ {\bf n} & =& ({\bf n}_0\cdot{\bf n}_L){\bf n}_L - \Bigl({\bf n}_0 - ({\bf n}_0\cdot{\bf n}_L){\bf n}_L\Bigr)\cos 2\xi + ({\bf n}_0\times{\bf n}_L)\,\sin 2\xi. \end{eqnarray} Since this is a global isospin rotation, the resulting field is another uniformly oriented solution with the same $\theta (x)$. For an axial rotation, $\xi = \eta$ and ${\bf n}_L = - {\bf n}_R$, in particular, if ${\bf n}_L, {\bf n}_R \;\bot\; {\bf n}_0$, one obtains \begin{eqnarray} \cos\theta(\tau) & = & \cos 2\xi\: \cos\theta_0(\tau), \nonumber \\ {\bf n}(\tau)\,\sin\theta(\tau) & = & {\bf n}_0\, \sin\theta_0(\tau) + {\bf n}_L\, \sin 2\xi\: \cos\theta_0(\tau). \end{eqnarray} The axial rotations turn a uniform solution into nonuniform solutions. We can actually show that our general nonuniform solution given by (16) and (18) is reproduced with a suitable choice of the rotation angle: \begin{equation} \tan 2\xi = a/b. \end{equation} Then (26) gives \begin{eqnarray} \cos\theta(\tau) & = & (b/\kappa)\cos\theta_0(\tau), \nonumber \\ \cos\,\beta(\tau) & = & {\bf n}\cdot{\bf n}_0 = \sin\theta_0(\tau)/\sin\theta(\tau). \end{eqnarray} The result in (18) for $\cos\beta$ is obtained by solving the above equations. In this way we are able to obtain the nonuniform solutions from uniform ones by the axial-vector rotations. In the boost-invariant 1+1 dimensional case, we have obtained in Sec.\ 2 all nonuniform solutions by solving explicitly the nonlinear differential equations. We have shown that they are all related by the axial rotations to the nonuniform solutions with the same energy density ${\cal E}(x)$. All possible solutions are exhausted in this way. \section{General Solution Without Boost Invariance} Even without boost invariance, it is straightforward to solve for the uniformly oriented solutions. Once we obtain a uniform solution, we can transform it into the nonuniform solutions by $SU(2)\times SU(2)$ rotations. The equations for $\theta(x)$ and ${\bf n}(x)$ from (5) to (8) in Sec.\ 2 are also valid in the boost-noninvariant case. We look for the uniform solution in which ${\bf n} = {\rm constant}$ so that $\partial_{\mu}{\bf n}=0$. In this case the equation of motion reduces simply to \begin{equation} \Box\theta = 0. \end{equation} It is convenient to use the space-time variables \begin{equation} \tau =\sqrt{t^2 - z^2},\;\;\; \eta =\frac{1}{2} {\rm ln}\Biggl(\frac{t+z}{t-z}\Biggr), \;\;\; {\bf x}_\bot. \nonumber \end{equation} The origin of space-time coordinates is identified with the collision point of the hadron collisions. The z-axis is chosen along the collision axis of the initial hadrons. Note that the meaning of variable $\tau$ is a little different from the 1+1 dimensional case. The surface of $\tau=0$ lies outside the light cone with respect to the space-time origin except for the exactly forward and backward directions. With these space-time variables, $\Box\theta = 0$ becomes \begin{equation} \frac{1}{\tau}\frac{\partial}{\partial\tau}\Biggl(\tau \frac{\partial\theta} {\partial\tau}\Biggr) - \frac{1}{\tau^2}\frac{\partial^2\theta}{\partial^2\eta}- \triangle_\bot\theta = 0. \end{equation} Since the differential equation is homogeneous, it can be solved by the method of separation of variables in the form \begin{equation} \theta(x) = T(\tau)H(\eta)X({\bf x_\bot})\:{\rm\theta}(\tau^2-{\bf x}_\bot^2). \end{equation} We solve for $\theta(x)$ inside the light cone, $\tau^2 - {\bf x}_\bot^2 > 0$. For the transverse direction ${\bf x_\bot}$, the general solutions are the Bessel and the Neumann functions. If we require that the solution be regular on the collision axis $\rho =|{\bf x_\bot}|= 0$, the Neumann functions are excluded. Note however that a singular behavior means an infinite oscillation toward $\rho = 0$, not an indefinite increase, in terms of the pion field $\mbox{\boldmath $\pi$} = f_{\pi}{\bf n}\, \sin\theta$. One may also require that $X({\bf x_\bot})$ should not increase indefinitely as $\rho \rightarrow\infty$. With these requirements, the parameter $\mu^2$ defined by $\triangle_\bot X = -\mu^2 X$ must be positive. We choose $\mu > 0$. $X({\bf x}_\bot)$ is expressed in terms of the Bessel functions of integer order: \begin{equation} X({\bf x_\bot}) = C_0J_0(\mu\rho) + \sum_{m=1}^\infty J_m(\mu\rho)(C_m \cos m\phi + D_m \sin m\phi), \end{equation} where $C_0$, $C_m$, and $D_m$ are the numerical coefficients to be determined by the boundary conditions. The magnitude of $\mu$ determines a transverse size of a DCC and therefore a spread of the $p_t$ distribution of the final pions. Since a DCC will have an extended size in the transverse direction, the value of $\mu$ is likely to be a fraction of $f_{\pi}$ or less. The spatial rapidity dependence $H(\eta)$ is simply solved \begin{equation} H(\eta) = \cosh \lambda\eta \;\;\; or \;\;\;\sinh \lambda\eta. \end{equation} The parameter $\lambda$ can be any complex number in general (If it is complex, one should take the real part of $\theta(x)$ at the very end). The special case $\lambda = 0$ leads to the boost-invariant solutions. For the approximately boost-invariant DCC configurations, the magnitude of $\lambda$ is much smaller than unity. The region of large values of $\eta$ corresponds to the forward and backward edges of DCC where energetic leading hadrons are moving outward while the small values of $\eta$ describe the cool central region. In this picture, it appears appropriate to choose $H(\eta)$ such that the pion density is higher at a larger $\eta$ than at a smaller $\eta$. We therefore choose $\lambda$ to be real (and positive) in the following. It should be emphasized however that our choice for a real $\lambda$ over purely imaginary or complex $\lambda$ is more for the convenience of the presentation. Given $X({\bf x_\bot})$ and $H(\eta)$, $T(\tau)$ obeys the differential equation \begin{equation} \frac{1}{\tau}\frac{\partial}{\partial\tau}\Biggl(\tau\frac{\partial T} {\partial \tau}\Biggr) + \Biggl(\mu^2 - \frac{\lambda^2}{\tau^2}\Biggr)T = 0. \end{equation} The solution is given by $ J_\lambda(\mu\tau)$ and/or $ N_\lambda(\mu\tau)$. The main difference between $J_\lambda(\mu\tau)$ and $N_\lambda(\mu\tau)$ is their different behavior as $\mu\tau$ approaches $0$. In the limit of a boost-invariance, $\lambda\rightarrow 0$, $J_\lambda (\mu\tau)$ approaches unity while $N_\lambda(\mu\tau) \rightarrow \ln (\mu\tau)$. If there is no transverse momentum, the latter approaches the uniform solution ($a = 0$, $\kappa = b$) described in Sec.\ 2, while the former coincides with the lowest energy solution given in (21). Putting all together, one obtains the uniform solution in a complete form: \begin{eqnarray} \theta(x){\bf n}_0 & = & \Bigl(a\:J_\lambda(\mu\tau) + b\:N_\lambda(\mu\tau)\Bigr)\nonumber \\ & & \times \Bigl(A\:\cosh\lambda\eta + B\:\sinh\lambda\eta\Bigr)\nonumber \\ & & \times J_m(\mu\rho)(C_m \cos m\phi + D_m \sin m\phi)\,{\bf n}_0\;{\rm \Theta}(\tau^2-{\bf x}_\bot^2), \end{eqnarray} where one may superpose these solutions in $\lambda$, $\mu$, and $m$. It is much harder to solve directly for nonuniform solutions when there is no boost-invariance constraint. Though some simple special solutions can be obtained by luck, finding all nonuniform solutions is a formidable task. In contrast, it is straightforward to perform $SU(2)\times SU(2)$ rotations on the uniform solutions. The rotation formulas (23) and (24) are most general and applicable to the boost-noninvariant case as well. Therefore the nonuniform solutions can be obtained by the chiral rotation from the uniform one in (36). An important question is whether we exhaust all nonuniform solutions by the axial rotations from the uniform solutions. In other words, are there any nonuniform solutions that cannot be rotated into a uniform one? If such a class of solutions exists, it would have some topological quantum number like a soliton. Note however that the solutions of our interest are time-dependent, and that their energies and actions are not necessarily finite. Unless these topologically nontrivial solutions exist, the uniform solution (36) and the $SU(2)\times SU(2)$ rotations on it exhaust all solutions. \section{Pion Charge Distribution} It has been predicted that the pions decaying from a DCC will show a distinct charge distribution when the charge ratio is plotted event by event in $f= N_{\pi^0}/(N_{\pi^0}+N_{\pi^\pm})$. The distribution \begin{equation} \frac{dP}{df} = \frac{1}{2\sqrt{f}} \label{dpdf} \end{equation} has been derived in two very different ways. The first derivation assumes that an isosinglet multipion state is created by the decay of a DCC \cite{horn}. All pions decaying from a given DCC are assumed to occupy an identical orbital state that is determined by the spatial configuration of DCC. The Bose statistics allows only one isosinglet $2N$-pion state: \begin{equation} |2N\pi\rangle = (2a_{+}^\dagger a_{-}^\dagger - a_0^\dagger a_0^\dagger)^N|0\rangle, \end{equation} where $a_{\pm,0}^\dagger$ are the creation operators of the pions in the same single orbital state. Making a binomial expansion of the right-hand side at large $N$, one obtains a simple rule $dP/df = 1/(2\sqrt{f})$. It is later pointed out that the relative phase between $a_{+}^\dagger a_{-}^\dagger$ and $a_0^\dagger a_0^\dagger$ is inessential to the final prediction of $dP/df$ \cite{kog}. The second derivation is based on a more intuitive picture in classical field theory. Assuming that the isospin orientation is uniform in space-time and that all isospin directions are equally probable, one obtains $dP/d\Omega = 1/4\pi$, where $\Omega$ is the solid angle for an isospin direction in isospin space. Since the $\pi^0$ fraction $f$ is proportional to the square of the third component of the pion field, ($f\propto\cos^2\beta$), one obtains again distribution (\ref{dpdf}). In this derivation, the interference effects are completely ignored. Let us examine whether or not this prediction remains valid for the nonuniform DCC's. In the first derivation, it is crucial that only one orbital state is available for pions and therefore the isosinglet state is unique: For two pions, the isosinglet is nothing but $(2a_{+}^\dagger a_{-}^\dagger - a_0^\dagger a_0^\dagger)|0\rangle$ by the Clebsch-Gordan coefficients. For four pions, the group theory alone would allow two isosinglets. One is to combine the ${\bf 0}_{2\pi}$ from ${\bf 1}_\pi\otimes{\bf 1}_\pi ={\bf 0}_{2\pi} + {\bf 1}_{2\pi}$ with the other ${\bf 0}_{2\pi}$ from ${\bf 1}_\pi\otimes{\bf 1}_\pi ={\bf 0}_{2\pi} +{\bf 1}_{2\pi}$. The other is to contract the ${\bf 1}_{2\pi}$ from ${\bf 1}_\pi\otimes{\bf 1}_\pi= {\bf 0}_{2\pi} + {\bf 1}_{2\pi}$ with the other ${\bf 1}_{2\pi}$ from ${\bf 1}_\pi\otimes{\bf 1}_\pi={\bf 0}_{2\pi}+ {\bf 1}_{2\pi}$. The Bose statistics forbids ${\bf 1}_{2\pi}$ for two identical pions in the same orbital state so that only ${\bf 0}_{2\pi}\otimes {\bf 0}_{2\pi} |0\rangle$ = $(2a_{+}^\dagger a_{-}^\dagger -a_0^\dagger a_0^\dagger)^2|0 \rangle$ is allowed. This argument goes through for any $2N$, leading to the $|2N\pi\rangle$ in (38). If there are more than one orbital states available, the four-pion singlet state would generally take the form \begin{equation} |4\pi\rangle = \Bigl(A({\bf 0}_{2\pi}\otimes{\bf 0}_{2\pi})+B({\bf 1}_{2\pi} \otimes{\bf 1}_{2\pi})\Bigr)|0\rangle, \end{equation} where the coefficients $A$ and $B$ are dynamics-dependent. There are increasingly many more isosinglets for $6\pi$'s, $8\pi$'s {\it etc}, as $N$ goes up. In the above example, the $A$-type term and the $B$-type term give quite different pion compositions: there is $\pi^0\pi^0\pi^0\pi^0$ in the $A$-type term, but no $\pi^0\pi^0\pi^0\pi^0$ in the $B$-type term. In order to obtain the distribution $dP/df = 1/(2 \sqrt{f})$, there must be only the $A$-type term and nothing else in the $2N\pi$ state ($N\rightarrow\infty$). One can construct explicitly the $4\pi$ state when the isovector field ${\bf n}(x)$ is nonuniform in space-time. Let us parametrize the direction of ${\bf n}(x)$ by the azimuthal and polar angles $\alpha(x)$ and $\beta(x)$ with respect to the isospin z-axis. To simplify our computation a little, we consider as an example a DCC whose isospin is nonuniform only in the polar direction $\beta$, but not in the azimuthal direction by choosing $\alpha = 0$. We shall use the representation $\mbox{\boldmath $\pi$}(x)= f_{\pi}{\bf n}(x)\sin \theta(x)$ instead of $\pi(x) = f_{\pi}{\bf n}(x)\theta(x)$ for the following discussion since the former automatically incorporates the periodicity of $\Sigma(x)$ in $\theta(x)\rightarrow\theta(x)\pm 2\pi$. The Cartesian isospin components of the pion field are \begin{eqnarray} \pi_1 & = & f_{\pi}\sin\theta(x)\;\sin\beta(x),\nonumber\\ \pi_2 & = & 0, \nonumber \\ \pi_3 & = & f_{\pi}\sin\theta(x)\;\cos\beta(x). \end{eqnarray} The DCC state is described by the quantum coherent state, up to an overall normalization \begin{equation} |DCC(\theta,\beta)\rangle = \exp\Bigl(a_1^\dagger(s_\theta s_\beta) + a_3^\dagger(s_\theta c_\beta)\Bigr)|0\rangle, \end{equation} where \begin{eqnarray} a_1^\dagger(s_\theta s_\beta) & = & \int\sqrt{2|{\bf k}|}\phi_{ss} ({\bf k})a_1^\dagger({\bf k}) d^3{\bf k},\nonumber \\ a_3^\dagger(s_\theta c_\beta) & = & \int\sqrt{2|{\bf k}|}\phi_{sc} ({\bf k})a_3^\dagger({\bf k}) d^3{\bf k}, \end{eqnarray} with $\phi_{ss}$ and $\phi_{sc}$ being the three-dimensional Fourier transforms of $f_{\pi}\sin\theta\sin\beta$ and $f_{\pi}\sin\theta\cos\beta$ respectively. Unlike $a_i^\dagger({\bf k})$, the operators $a_1^\dagger(s_\theta s_\beta)$ and $a_3^\dagger(s_\theta c_\beta)$ are not canonically normalized, but the normalization is irrelevant to the isospin structure. The $|N\pi\rangle$ projection of the DCC state is \begin{equation} |N\pi(\theta(x)\,\beta(x))\rangle =\frac{1}{N!}\Bigl(a_1^\dagger(s_\theta\,s_\beta) + a_3^\dagger (s_\theta\,c_\beta)\Bigr)^N|0\rangle. \end{equation} Under the assumption that the DCC's appear in the intermediate state with $I=0$ and the production processes conserve isospin, if one DCC can be produced, all other DCC's that are related to it by the isospin rotations can be produced with an equal probability. The isosinglet DCC state can be constructed from the state in (41) by integrating out the Euler angles over the entire isospin space. The $4\pi$ state of an isosinglet DCC is obtained by averaging $|4\pi(\theta(x)\, \beta(x))\rangle$ over isospin space. The computation is straightforward though a little tedious. Up to an overall normalization, the result is \begin{equation} |4\pi(I=0)\rangle = \Biggl(\Bigl(\Bigl|{\bf a}^\dagger(s_\theta c_\beta)\Bigr|^2 +\Bigl|{\bf a}^\dagger(s_\theta s_\beta)\Bigr|^2\Bigr)^2 -4\Bigl|{\bf a}^\dagger(s_\theta c_\beta)\times {\bf a}^\dagger(s_\theta s_\beta)\Bigr|^2\Biggr)|0\rangle, \end{equation} where $|{\bf a}^\dagger|^2 = 2a_{+}^\dagger a_{-}^\dagger - a_0^\dagger a_0^\dagger$. The first and second terms in the right-hand side are the $A$-type terms in (39), while the last term is the $B$-type term. As we anticipate, the isosinglet $4\pi$ state of the nonuniform DCC is no longer of the form postulated in (38). For a uniform DCC, that is, $\beta(x)\rightarrow\,{\rm constant}$, ${\bf a}^\dagger(s_\theta c_\beta)$ and ${\bf a}^\dagger(s_\theta s_\beta)$ are identical up to a factor (in the 1+1 boost-invariant solution in Sec.\ 2, $\beta(x)$ is so defined that $\beta(x)\rightarrow 0$, namely ${\bf a}^\dagger\rightarrow 0$, in the uniform limit). Therefore, the $B$-type term cannot exist for the uniform DCC's. Our construction of the isosinglet $4\pi$ state and the existence of the $B$-type terms cast a serious doubt on the distribution for the nonuniform DCC's. Alternatively, let us study the problem by assuming that $|{\rm DCC}(\theta(x),\beta(x))\rangle$ with different $\theta(x)$ and $\beta(x)$ do not have the quantum interference with each other. It is in accordance with the classical field picture. For a large number of pions, ignoring the interference may be justified. The momentum spectrum of pion quanta ($i=1,2,3$) decaying from a classical field is given \cite{hen} \begin{equation} (2\pi)^3\frac{dN_i}{d^3{\bf k}} = \frac{|\tilde{\rho}_i({\bf k},|{\bf k}|)|^2}{2|{\bf k}|}, \label{dn} \end{equation} where $\tilde{\rho}_i({\bf k},|{\bf k}|)$ is the four-dimensional on-mass-shell Fourier transform of the pion source function $\rho_i({\bf x},t)$ defined by $\Box \mbox{\boldmath $\pi$}(x)=\mbox{\boldmath $\rho$}(x)$: \begin{equation} \tilde{\rho}_i({\bf k}, |{\bf k}|) = \int \rho_i({\bf x}, t)\, e^{-i{\bf k}\cdot{\bf x} +i|{\bf k}|t}d{\bf x}dt. \end{equation} It is convenient to perform the space-time integral using variables $\tau,\;\eta$ and ${\bf x}_\bot$ for which $d{\bf x}dt = \tau d\tau d\eta d{\bf x}_\bot$, and \begin{equation} E= |{\bf k}|, \;\;\; y =\frac{1}{2}{\rm ln} \Biggl(\frac{E+k_{\|}}{E-k_{\|}}\Biggr), \;\;\;{\bf k}_\bot ,\label{m} \end{equation} for the momentum variables. (\ref{dn}) becomes \begin{equation} (2\pi)^3\frac{dN_i}{dy\,d^2{\bf k}_\bot} = \Bigl|\int \rho_i(\tau,\,\eta,\,{\bf x}_\bot)\;e^{i|{\bf k}_\bot|\tau \,\cosh(\eta - y) -i{\bf k}_\bot\cdot{\bf x}_\bot } \tau\,d\tau\,d\eta\,d^2{\bf x}_\bot\Bigr|^2. \end{equation} If a DCC is boost-invariant along the collision axis, $\rho(\tau,\eta,{\bf x}_\bot)$ does not depend on $\eta$. In this case, $\eta$ is integrated out and the energy spectrum $dN_i/dy\,d^2{\bf k}_\bot$ is independent of the rapidity variable $y$, as it is well known. Let us look into the isospin structure of the Fourier transform of the source that enters the right-hand side of (48). After the space-time integration is performed, the isovector $\tilde{\mbox{\boldmath $\rho$}}$ is generally of the form \begin{equation} \tilde{\mbox{\boldmath $\rho$}}({\bf k}, |{\bf k}|) = F(y,{\bf k}_\bot)\mbox{\boldmath $e$}({\bf k})\; , \end{equation} where $F(y,{\bf k}_\bot)$ is an isoscalar, Lorentz-scalar function of ${\bf k}$ and of whatever parameters that characterize a DCC; $\mbox{\boldmath $e$}({\bf k})$ is a unit vector in isospin space. The pion spectrum is simply \begin{equation} (2\pi)^3\frac{dN_i}{dy\,d^2{\bf k}_\bot} = |F(y,{\bf k}_\bot) e_i({\bf k})|^2. \end{equation} For each {\em fixed} ${\bf k}$, one may repeat the classical field derivation and reproduce \begin{equation} \frac{dP}{df({\bf k})} = \frac{1}{2\sqrt{f({\bf k})}}.\label{dpdfk} \end{equation} However, it is clear that the pion number $N_i$ no longer obeys distribution (\ref{dpdf}) when the momentum ${\bf k}$ is integrated over. To illustrate this point, consider a toy DCC for which $\mbox{\boldmath $e$}({\bf k})$ points to one direction for a half of the range of rapidity $y$ and to another direction perpendicular to it for the other half of $y$. Such a DCC is not one of the solutions that we have obtained, but it serves to make a point. Since there is no way to align the two ${\bf e}({\bf k})$'s to the same direction by isospin rotation, there are no DCC's in this isospin family that emit only $\pi^0$, even though all directions are equally probable in isospin space. For a family of nonuniform DCC's, $dP/df$ is zero at $f=0$ ({\em Centauro}) and at $f=1$({\em anti-Centauro}), and tends to bulge in the central region of $f$, unlike that for a family of uniform DCC's. Only if the uniform DCC's dominate over the nonuniform ones, can distribution (\ref{dpdf}) hold approximately. The abundance of the uniform DCC's has a measure zero relative to that of the nonuniform DCC's in the phase space of the rotation angles. Unless the production of the nonuniform DCC's by the initial hadrons is strongly suppressed for some dynamical reason, $dP/df= 1/(2\sqrt{f})$ cannot hold even approximately. The spectacular Centauro and anti-Centauro events will be far rarer than our naive expectation based on the uniform DCC's. However, there may be a chance to observe distribution (\ref{dpdfk}) by selecting pions of the same $y$ and ${\bf k}_\bot$ within small uncertainties. A special case is a boost-invariant DCC in $1+1$ dimensions where $\mbox{\boldmath $\rho$}(x)$ is $\eta$-independent so that $\tilde{\mbox{\boldmath $\rho$}}({\bf k})$ is $y$-independent. In this case, $\mbox{\boldmath $e$}({\bf k})$ becomes a constant vector independent of ${\bf k}$ and distribution (\ref{dpdf}) follows even for the nonuniform DCC's. One the other hand, for the $4\pi$ state that we studied in this Section, we see nothing special about the boost-invariant nonuniform case with ${\bf k}_\bot = 0$ from the general boost-noninvariant case. Do two derivations contradict with each other? It is difficult to make a connection between the two arguments. In analyzing the $|N\pi(I=0)\rangle$ state, the interference between different DCC's is essential while in the classical field analysis, each DCC state is not an eigenstate of isospin and the interference from different DCC's is completely discarded. Though the both methods have led to the same $dP/df$ distribution, it is not clear how much similar or mutually compatible their physical pictures are. With this unsolved uncertainty, we state our conclusion in a less assertive way: if we follow the isospin analysis of $|2N\pi\rangle$, we see no mechanism that leads to distribution (\ref{dpdf}) for the nonuniform DCC's. If we argue instead in the classical field picture, the distribution does not hold except for the boost-invariant DCC with zero ${\bf k}_\bot$. However, distribution (\ref{dpdf}) should hold for pions which are selected from a small segment of rapidity $y$ and transverse momentum ${\bf k}_\bot$. \section*{Acknowledgment} We wish to thank J.D.\ Bjorken for giving us his notes on his solutions and many useful discussions. This work was supported in part by the Director, Office of Energy Research, Office of High Energy and Nuclear Physics, Division of High Energy Physics of the U.S. Department of Energy under contract DE-AC03-76SF00098 and in part by the U.S. National Science Foundation under grant PHY-90-21139. One of us (Z.H.) acknowledges the support from the Natural Sciences and Engineering Research Council of Canada. \newpage
\section{#1}} \renewcommand{\theequation}{\thesection.\arabic{equation}} \newcommand{\app}[1]{\setcounter{section}{0} \setcounter{equation}{0} \renewcommand{\thesection} {\Alph{section}} \section{#1}} \def\noindent{\noindent} \def\begin{array}{rcl}{\begin{array}{rcl}} \def\end{array}{\end{array}} \def\rangle{\rangle} \def\langle{\langle} \newcommand{\begin{equation}}{\begin{equation}} \newcommand{\begin{eqnarray}}{\begin{eqnarray}} \newcommand{\end{equation}}{\end{equation}} \newcommand{\end{eqnarray}}{\end{eqnarray}} \def{\scriptscriptstyle {\cal F}_{12}}{{\scriptscriptstyle {\cal F}_{12}}} \def{\scriptscriptstyle {\cal F}_{12\ldots n}}{{\scriptscriptstyle {\cal F}_{12\ldots n}}} \def{\cal F}_{12}{{\cal F}_{12}} \def{\cal F}_{12\ldots n}{{\cal F}_{12\ldots n}} \def{\cal H}{{\cal H}} \def\otimes{\otimes} \def\mbox{id}{\mbox{id}} \def\mbox{\it i.e.\/ }{\mbox{\it i.e.\/ }} \def\mbox{\it e.g.\/ }{\mbox{\it e.g.\/ }} \def{\cal C}{{\cal C}} \def1_{\cal A}{1_{\cal A}} \def1_{\cal U}{1_{\cal U}} \def\Upsilon{\Upsilon} \def\mbox{\bf g}{\mbox{\bf g}} \def\mbox{$U_{q}{\/\mbox{\bf g}}$}{\mbox{$U_{q}{\/\mbox{\bf g}}$}} \def\mbox{Fun$(G_{q})$}{\mbox{Fun$(G_{q})$}} \newcommand{\triangleright}{\triangleright} \newcommand{{\,\stackrel{s}{\triangleright}\,}}{{\,\stackrel{s}{\triangleright}\,}} \def\mbox{$\times \!\rule{0.3pt}{1.1ex}\,$}{\mbox{$\times \!\rule{0.3pt}{1.1ex}\,$}} \def{\mbox{$\cal A$} \cross \mbox{$\cal U$}}{{\mbox{$\cal A$} \mbox{$\times \!\rule{0.3pt}{1.1ex}\,$} \mbox{$\cal U$}}} \def\mbox{$\cal R$}{\mbox{$\cal R$}} \newcommand{\mbox{$\cal Y$}}{\mbox{$\cal Y$}} \newcommand{\mbox{$\cal Z$}}{\mbox{$\cal Z$}} \def\mbox{$\cal A$}{\mbox{$\cal A$}} \def\mbox{$\cal U$}{\mbox{$\cal U$}} \def\mbox{$\cal F$}{\mbox{$\cal F$}} \def\mbox{$\cal V$}{\mbox{$\cal V$}} \def\mbox{$\cal O$}{\mbox{$\cal O$}} \newcommand{\Delta _{\cal A}}{\Delta _{\cal A}} \newcommand{{}_{\cal A}\Delta }{{}_{\cal A}\Delta } \newcommand{{}_{\cal U}\Delta }{{}_{\cal U}\Delta } \newcommand{\Delta_{\cal U}}{\Delta_{\cal U}} \newcommand{\mbox{d}}{\mbox{d}} \def\hspace*{9mm}{\hspace*{9mm}} \def\hspace{3mm}{\hspace{3mm}} \newcommand{\stackrel{\mbox{\scriptsize ad}}{\triangleright}}{\stackrel{\mbox{\scriptsize ad}}{\triangleright}} \def{\cal I}_{\rm left}{{\cal I}_{\rm left}} \def{\Gamma}_{\rm left}{{\Gamma}_{\rm left}} \def\mbox{\boldmath $i$}{\mbox{\boldmath $i$}} \def\Ix#1{\mbox{\boldmath $i$}_{\chi_#1}} \def\hbox{\large\it \pounds}{\hbox{\large\it \pounds}} \def\Lix#1{\hbox{\large\it \pounds}_{\chi_#1}} \def\Lio#1#2{\hbox{\large\it \pounds}_{O_#1{}^#2}} \def\mbox{\bf d}{\mbox{\bf d}} \newcommand{\mbox{\bf D}}{\mbox{\bf D}} \newcommand{\mbox{$\omega$}}{\mbox{$\omega$}} \newcommand{\alpha}{\alpha} \newcommand{\epsilon}{\epsilon} \newcommand{\mbox{Fun({\bf M}$_{q}$)}}{\mbox{Fun({\bf M}$_{q}$)}} \newcommand{\mbox{${\cal T}_q(\mbox{\bf M}_q)$}}{\mbox{${\cal T}_q(\mbox{\bf M}_q)$}} \newcommand{\mbox{${\cal T}_q({G}_q)$}}{\mbox{${\cal T}_q({G}_q)$}} \def\mbox{${\cal T}_q$}{\mbox{${\cal T}_q$}} \def\mbox{\boldmath$\delta $}{\mbox{\boldmath$\delta $}} \newcommand{\stackrel{\mbox{\scriptsize ad}}{\triangleleft}}{\stackrel{\mbox{\scriptsize ad}}{\triangleleft}} \newcommand{{\bf C}}{{\bf C}} \newcommand{{\bf R}}{{\bf R}} \newcommand{{\bf Z}}{{\bf Z}} \newcommand{{\bf N}}{{\bf N}} \begin{document} \begin{titlepage} \begin{center} August 1995 \hfill LMU-TPW 95-10\\ \mbox{} \hfill (revised, Feb.\ 1996)\\ \vskip.6in {\Large \bf Identical Particles and Quantum Symmetries} \vskip.4in Gaetano Fiore* and Peter Schupp \vskip.25in {\em Sektion Physik der Ludwig-Maximilians-Universit\"at M\"unchen\\ Theoretische Physik --- Lehrstuhl Professor Wess\\ Theresienstra\ss e 37, 80333 M\"unchen\\ Federal Republic of Germany} \end{center} \vskip1in \begin{abstract} We propose a solution to the problem of compatibility of Bose-Fermi statistics with symmetry transformations implemented by compact quantum groups of Drinfel'd type. We use unitary transformations to conjugate multi-particle symmetry postulates, so as to obtain a twisted realization of the symmetric groups $S_n$. \end{abstract} \vfill \noindent \hrule \vskip.2cm \noindent{\footnotesize *A.~v.~Humboldt-fellow\hfill {\it e-mail: } Fiore, Schupp \ @ \ ls-wess.physik.uni-muenchen.de} \end{titlepage} \newpage \setcounter{page}{1} \sect{Introduction} Quantum groups \cite{dr,ji,frt} have received much attention in recent years as candidates for generalized symmetry transformations in physics. Among other applications, they look promising in relation to generalized space-time\footnote{These are symmetries of a proposed non-commutative structure of space-time \cite{wess}.} and/or internal symmetries in Quantum Field Theory. One way to approach QFT consists first in finding a consistent procedure to implement quantum group transformations in Quantum Mechanics with a finite number of particles, then to pass to QFT through second quantization. Various models describing systems of one particle (see e.g. ref. \cite{wess1,wess2,fio,fioeu,wei}) or a finite number of {\em distinct} particles consistently transforming under the action of a quantum group have been constructed so far; as known, the quantum group coproduct plays a specific role in extending quantum group transformations from one-particle to multi-particle systems. In this article we would like to study whether the notions of identical particles and quantum group transformations are compatible in quantum mechanics (in first quantization). The setting that we have in mind is a quantum mechanical system transforming under generalized (symmetry) transformations realized by some $*$-Hopf algebra $H$ \footnote{The transformations may correspond to a symmetry either in the sense that they leave the {\it dynamics} of the particular system under consideration invariant (e.g. rotation symmetry of its Hamiltonian), and therefore are associated to conservation laws for the latter; or in the sense that they leave the {\it form} of the physical description of {\it any} system invariant (covariance of the physical description), as it happens e.g. with the Poincar\'e transformations in Special Relativity.} (in particular, a $*$-quantum group \cite{dr}). In order that a system of $n$ bosons/fermions transforms under the action of $H$ its Hilbert space of states should carry both a representation of the symmetric group $S_n$ and of $H$. In the case that the $H$ is quantum group, one might expect that this is impossible. We recall that in the standard quantum mechanical formalism the elements of $S_n$ are realized as ordinary permutation operators. On the other hand, in the Hopf algebra formalism the action of $H$ on a multiparticle system is defined through the coproduct $\Delta$. Given a representation $\rho$ of $H$ on a ``one-particle'' Hilbert space ${\cal H}$, and considering (for simplicity) the case of two particles, the action of $H$ on ${\cal H} \otimes {\cal H}$ is defined through $(\rho\otimes\rho)\circ \Delta$. In the case that $H$ is cocommutative (\mbox{\it e.g.\/ } $H=U(su(2))$), the coproduct takes the form $\Delta(X_i)= X_i\otimes 1 + 1\otimes X_i$ on all the generators $X_i$ (in the case $H=U(su(2))$ this expresses the classical addition law of angular momentum); therefore the above action preserves the symmetric and antisymmetric subspaces $({\cal H} \otimes {\cal H})_\pm$ defined by $P_{12} ({\cal H} \otimes {\cal H})_\pm = \pm ({\cal H} \otimes {\cal H})_\pm$ respectively ($P_{12}$ denotes the permutation operator). When $H$ is not cocommutative, \mbox{\it e.g.\/ } it is a quantum group, $\Delta$ is no more symmetric under the action of $P_{12}$, so that the above action mixes $({\cal H} \otimes {\cal H})_+$ and $({\cal H} \otimes {\cal H})_-$. Therefore, fermions and bosons in the ordinary sense seem impossible, and it is natural to speculate that in the quantum group context some new (or ``$q$-") statistics is necessary or even that the notion of identical particles must be abandoned. Even if $H$ is just a {\em slight\/} deformation of a co-commutative Hopf algebra (e.g. an ordinary Lie group) a new statistics would result into a drastic discontinuity of the number of allowed states of the multi-particle system in the limit of vanishing deformation parameter ($\ln q$ in the $H=\mbox{$U_{q}{\/\mbox{\bf g}}$}$-case): in fact, elementary particles cannot be ``almost identical'', they can only be either identical or different. However, such a discontinuity appears physically unacceptable if we think of $H$ as a slight modification of some experimentally well-established symmetry of elementary particle physics. A previously suggested ``quick fix'' of the problem is the naive symmetrization of coproducts ---this approach will however destroy any true quantum symmetry. It is also important to realize that it is not enough to make sense of $\Delta(H), \Delta^2(H),$ {\em etc}. The spaces of multi-particle operators have to be larger than that to be in one-to-one correspondence with their classical (symmetrized) counterparts. In the $H=U_q(su(2))$ case, for instance, we would like to construct the q-analog of the (classically) symmetric operators $X_i\otimes X_i$, which are not the coproduct of anything. In this work we want to show that a solution to the problem is a modification of our notions of symmetry and anti-symmetry associated to bosons and fermions. The point is that ordinary permutations are not the only possible realization of elements of the abstract group $S_n$; an alternative one can be obtained by applying some unitary transformation $F_{12\ldots n}$ to the permutators (see section \ref{twistmult}). The question (see section \ref{qsym}) is therefore whether for any number of particles $n$ there exists some $F_{12\ldots n}$ (the ``twist") such that the corresponding realization of $S_n$ is compatible with the action of $H$. Due to some theorems by Drinfel'd, this turns out to be the case {\em at least\/} if $H = \mbox{$U_{q}{\/\mbox{\bf g}}$}$ \cite{dr,ji,frt} is one of the standard quantum groups associated to the compact\footnote{For U${}_q$({\bf g}) this requires $q \in {\bf R}$. To study the problem in the case of $q$ on the unit circle the reader should consult \cite{mascho} for the structure of weak quasi-Hopf algebras.} simple Lie algebras $\mbox{\bf g}$ of the classical series---the case of $U_q \mbox{su}(2)$ will be studied in some detail in section \ref{example}---or if $H$ is a triangular Hopf algebra arising from the quantization of a solution of the classical Yang-Baxter equation \cite{drinf,Taktadjan} or from a twist of type \cite{resh} as {\em e.g.}\/ studied in \cite{luk}\footnote{Twisted coproducts are here interpreted as clustered 2-particle states.}. The precise criterion is that $H$ must be the twist of a co-commutative (quasi-)Hopf algebra \cite{Drinfeld}; in either case we also need the existence of a $*$-conjuga\-tion. In the case where $H$ is a quasi-triangular Hopf algebra one might have expected to see anyons arise as a consequence of the braid group character of $\mbox{$\cal R$}$; however, in our formulation this does not happen: The statistics parameter is not modified---bosons stay bosons\footnote{See \cite{voza,nelson} and the extended list of references therein for a discussion of this point in the context of $q$-deformed oscillators.}, fermions stay fermions, and anyons (though not studied explicitly) stay anyons. The extreme case of $q = -1$ is especially instructive in this context \cite{zachos,zhang}. Let us ask now how in the context of identical particles the existence of quantum group symmetries of the above kind could manifest itself experimentally: The dynamical evolution of a system of $n$ identical particles will contain new physics only if we adopt an Hamiltonian which is {\em natural\/} to the twisted picture. One can always obtain a Hamiltonian {\em consistent\/} with twisted symmetrization postulates by a unitary transformation (through $F_{12\ldots n}$) on a Hamiltonian corresponding to some undeformed model, however, such a Hamiltonian will in general be of a very complicated, {\em i.e.\/} unnatural form. In section \ref{onemany} we will analyze a scattering experiment to see how the twist will manifest itself in the transformation of the initial and final data (which is essentially the tensor product of one-particle states) into the equivalent twisted (anti-)symmetrized states upon which the evolution operator describing the scattering acts. The $F_{12\ldots n}$ can again be absorbed in a redefined Hamiltonian, so that an experiment cannot decide whether we are in the twisted picture or not. (It is just a change of base.) We can only tell what picture is more natural. The main message is then that the twisted picture can be consistently introduced. (Contrary to expectation, there are no problems with statistics.) The twisted picture may lead to the development of models that one would probably not think of otherwise. One would expect to see {\em direct\/} consequences of the twists only in particle creation and annihilation processes; this however belongs to the realm of quantum field theory and will be treated elsewhere. For readers not familiar with the notion of Hopf algebras, we give a very brief introduction to the subject in section \ref{appendix}. After completion of this work we became aware of the very interesting paper in Ref. \cite{wopu}, which gives a quantization scheme for fields transforming covariantly under $SU_q(N)$. In a future work we will compare our results with the ones therein while considering the issue of second quantization. \sect{Twisted Multi-Particle Description} \label{twistmult} Let us forget the issue of quantum symmetry and hence the coproduct for the moment, and just consider pure quantum mechanics for identical particles. Consider a one-particle system, denote by ${\cal H}$ the Hilbert space of its states, and by $\mbox{$\cal A$}$ the $*$-algebra of observables acting on ${\cal H}$. $n$-particle states and $n$-particle operators will live in as yet to be determined subspaces of ${\cal H}^{\otimes n}$ and $\mbox{$\cal A$}^{\otimes n}$ respectively. Let us consider states of two identical particles. The corresponding state vector $|\psi^{(2)}\rangle$ will be some element of the tensor product of two copies of the one-particle Hilbert space ${\cal H}$. Let $ P_{12}$ be the permutation operator on ${\cal H} \otimes {\cal H}$:\ $ P_{12}(|a\rangle \otimes |b\rangle) \equiv |b\rangle \otimes |a\rangle$. (In the sequel we will also use the symbol $\tau$ to denote the abstract permutation map of two tensor factors, $ \tau(a \otimes b) \equiv b \otimes a$.) The fact that we are dealing with identical particles manifests itself in the properties of state vectors under permutation: \begin{equation} P_{12}|\psi^{(2)}\rangle = e^{i \nu}|\psi^{(2)}\rangle,\label{stat} \end{equation} where $\nu = 0$ for Bose-statistics and $\nu = \pi$ for Fermi-statistics. For the corresponding expectation value of an arbitrary operator ${\cal O} \in \mbox{$\cal A$} \otimes \mbox{$\cal A$}$ we then find \begin{equation} \langle\psi^{(2)}|{\cal O}|\psi^{(2)}\rangle = \langle\psi^{(2)}| P_{12}\: {\cal O}\: P_{12}|\psi^{(2)}\rangle \end{equation} because the phases $e^{-i \nu}$ and $e^{i \nu}$ from the bra and the ket cancel. This means that the operators $\cal O$ and $ \tau({\cal O}) \equiv P_{12} {\cal O} P_{12}$ are members of the same equivalence class as far as expectation values go. One particular representative of each such equivalence class is the symmetrized operator \begin{equation} {1 \over 2}({\cal O} + \tau({\cal O}) ) \in (\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+. \end{equation} It plays a special role because it preserves the two-particle Hilbert spaces for any statistic (\ref{stat}), as we will recall below. We can hence avoid redundant operators by restricting $\mbox{$\cal A$} \otimes \mbox{$\cal A$}$ to the sub-algebra \begin{equation} (\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+ := \{ a \in \mbox{$\cal A$} \otimes \mbox{$\cal A$} : [P_{12}, a ] = 0 \} \label{symop} \end{equation} (note that $[P_{12}, a ] = 0 \Leftrightarrow \tau(a) = a$). In this article we will show how to find an analog of $(\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+$ compatible with quantum group transformations. We summarize the relevant equations characterizing a system of two bosons or fermions: \begin{eqnarray} && P_{12}|u\rangle_\pm = \pm|u\rangle_\pm \hspace*{9mm} \mbox{for} \hspace{3mm} |u\rangle_{\pm} \in ({\cal H} \otimes {\cal H})_\pm \label{fb}\label{from}\\ && a:({\cal H} \otimes {\cal H})_\pm \rightarrow ({\cal H} \otimes {\cal H})_\pm \hspace*{9mm} \mbox{for} \hspace{3mm} a \in (\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+ \label{ahh}\\ && *_2: (\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+\rightarrow (\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+, \hspace*{9mm}\mbox{where}\hspace{3mm} *_2 \equiv * \otimes * .\label{to} \end{eqnarray} Equation (\ref{fb}) defines bosonic $(+)$ and fermionic $(-)$ states as in (\ref{stat}). Equation (\ref{ahh}) follows from $[P_{12}, (\mbox{$\cal A$}\otimes\mbox{$\cal A$})_+] = 0$ and shows that symmetrized operators transform boson states into bosons states and fermion states into fermion states. Similar statements as given here for two particles obviously apply also to states of 3 and more identical particles and to other statistics (anyons). Can one also describe in a non-standard way the system of $n$ identical particles, using what we know for one particle, so that the description is perfectly consistent from the physical viewpoint? Let us concentrate on two-particle systems for the moment: For a unitary and in general not symmetric operator $F_{12} \in \mbox{$\cal A$} \otimes \mbox{$\cal A$}$, $F_{12}^{*_2} = F_{12}^{-1}$ where $*_2 = * \otimes *$, we define \begin{eqnarray} ({\cal H} \otimes {\cal H})_{\pm}^{{F_{12}}} & := & F_{12} ({\cal H} \otimes {\cal H})_\pm\\ P_{12}^{F_{12}}& := & F_{12} P_{12} F_{12}^{-1} \\ (\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+^{F_{12}}& := & F_{12}(\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+ F_{12}^{-1} \end{eqnarray} where $(\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+$ is as given above. We then find in complete analogy to equations (\ref{from} -- \ref{to}) \begin{eqnarray} && P_{12}^{F_{12}}|u\rangle_\pm = \pm|u\rangle_\pm \hspace*{9mm} \mbox{for} \hspace{3mm} |u\rangle_\pm \in ({\cal H} \otimes{\cal H})_\pm^{F_{12}}\\ && a:({\cal H} \otimes {\cal H})^{F_{12}}_\pm \rightarrow ({\cal H} \otimes {\cal H})^ {F_{12}}_\pm \hspace*{9mm} \mbox{for}\hspace{3mm} a \in (\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+^{F_{12}}\label{aHH}\\ && *_2: (\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+^{F_{12}}\rightarrow (\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+^{F_{12}} \end{eqnarray} and $a^{F_{12}}:= F_{12} a F_{12}^{-1}$ is hermitean iff $a$ is. Equation (\ref{aHH}) follows from \begin{equation} [P_{12}^{F_{12}}, (\mbox{$\cal A$}\otimes\mbox{$\cal A$})_+^{F_{12}}] = 0. \label{commop} \end{equation} In general, $({\cal H} \otimes {\cal H})^{F_{12}}_\pm$ will not be (anti-)symmetric, nor will $(\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+^{F_{12}}$ be symmetric. Can we still interpret $({\cal H} \otimes {\cal H})_\pm^{F_{12}}$ as the Hilbert space of states of the system of two bosons or fermions of equal type and $(\mbox{$\cal A$} \otimes \mbox{$\cal A$})_+^{F_{12}}$ as the corresponding $*$-algebra of observables? We can. In fact, we have just conjugated the standard description of the 2-particle system through $F_{12}$ into a unitary equivalent one. (This agrees with the general viewpoint put forward in \cite{eng,engnote}. See also next section for a discussion from the physical point of view.) Obviously the idea of conjugation can be generalized to a system of $n$ identical particles: Let $F_{12\ldots n} \in \mbox{$\cal A$}^{\otimes n}$ be unitary, \mbox{\it i.e.\/ } $(F_{12\ldots n})^{*_n} = (F_{12\ldots n})^{-1}$, where $*_n := *^{\otimes n}$, and define \begin{eqnarray} ({\cal H} \otimes\ldots\otimes {\cal H})_\pm^{F_{12\ldots n}}& := & F_{12\ldots n} ({\cal H} \otimes\ldots\otimes {\cal H})_\pm \label{def1}\\ P_{12}^{F_{12\ldots n}}& := & F_{12\ldots n} P_{12} (F_{12\ldots n})^{-1}\\ & \vdots & \nonumber \\ P_{n-1,n}^{F_{12\ldots n}}& := & F_{12\ldots n} P_{n-1,n} (F_{12\ldots n})^{-1}\\ (\mbox{$\cal A$} \otimes\ldots\otimes\mbox{$\cal A$})_+^{F_{12\ldots n}}& := & F_{12\ldots n} (\mbox{$\cal A$}\otimes\ldots\otimes\mbox{$\cal A$})_+(F_{12\ldots n})^{-1} \label{twistop} \end{eqnarray} where $$ (\mbox{$\cal A$}\otimes\ldots\otimes\mbox{$\cal A$})_+ := \{ a \in \mbox{$\cal A$}\otimes\ldots\otimes\mbox{$\cal A$} : [P_{i,i+1} , a] =0 , i=1,\ldots n-1\}, $$ and $P_{i,i+1}$ is the permutator of the $i^{th},(i\!+\!1)^{th}$ tensor factors. Then \begin{eqnarray} &&P_{i,i+1}^{F_{12\ldots n}}|u\rangle_\pm = \pm |u\rangle_\pm \hspace*{9mm}\mbox{for}\hspace{3mm} |u\rangle_\pm \in ({\cal H} \otimes\ldots\otimes {\cal H})_\pm^{F_{12\ldots n}} \label{astate}\\ &&a: ({\cal H} \otimes\ldots\otimes {\cal H})_\pm^{F_{12\ldots n}}\rightarrow ({\cal H} \otimes\ldots\otimes {\cal H})_\pm^{F_{12\ldots n}} \label{aHdH}\\ && \hspace*{9mm}\z\hspace*{9mm}\z\hspace*{9mm}\mbox{for}\hspace{3mm} a \in (\mbox{$\cal A$} \otimes\ldots\otimes\mbox{$\cal A$})_+^{F_{12\ldots n}}\\ &&*_n: (\mbox{$\cal A$} \otimes \ldots \otimes\mbox{$\cal A$})_+^{F_{12\ldots n}}\rightarrow (\mbox{$\cal A$} \otimes \ldots \otimes\mbox{$\cal A$})_ +^{F_{12\ldots n}}. \end{eqnarray} Equation (\ref{aHdH}) follows from \begin{equation} [P_{i,i+1}^{F_{12\ldots n}}, (\mbox{$\cal A$}\otimes\ldots\otimes\mbox{$\cal A$})_+^ {F_{12\ldots n}}] = 0. \label{asop} \end{equation} Note that in eqs. (\ref{astate}) to (\ref{asop}) the twist $F_{12\ldots n}$ does not explicitly appear any more; these equations give an {\it intrinsic} characterization of the twisted multi-particle description, involving only the operators $P_{i,i+1}^{F_{12\ldots n}}$. By construction $P_{i,i+1}^{F_{12\ldots n}}$ is hermitean, its square is the identity and (consequently) has only eigenvalues $\pm 1$; moreover, the degeneracy of these eigenvalues is the same as in the untwisted case. The operators $P_{i,i+1}^{F_{12\ldots n}}$ give a realization of the group $S_n$ of permutation of $n$ objects, because they satisfy the same algebraic relations as the ordinary permutators $P_{i,i+1}$; correspondingly, $(\mbox{$\cal A$}\otimes\ldots\otimes\mbox{$\cal A$})_+^{F_{12\ldots n}}$, $({\cal H} \otimes\ldots\otimes{\cal H})_\pm^{F_{12\ldots n}}$ carries irreducible representations of $S_n$. Viceversa, one could easily prove that, given operators satisfying these conditions, one can find a unitary $F_{12\ldots n}$ such that equations (\ref{def1}) to (\ref{twistop}) hold. It will turn out that, even though the twists which are relevant for the quantum symmetry issue are very hard to compute, the $P_{i,i+1}^{F_{12\ldots n}}$ are much less so; see section~\ref{example}. \noindent {\em Remark:} If we replace the nilpotent $P_{12}$ by some braid group generator one could also conjugacy transform anyons. \sect{Identical Versus Distinct Particles } \label{onemany} In some situations particles of the same kind can be equivalently treated as {\it identical} or {\it distinct}, and there exists a precise correspondence between these two descriptions. The twist $F$ directly enters the rule governing this correspondence while in the twisted postulates (\ref{astate}) to (\ref{asop}) (intrinsic formulation) it appears only hidden in the $P^F$ (together with its inverse). Transforming one kind of description into the other one is often needed for practical purposes, as we illustrate by the following example. Consider a gedanken experiment of a scattering of two identical particles. One can distinguish three stages. In the initial stage, the two particles are far apart and are assumed to be prepared in two separate one-particle normalized states $|\psi_1\rangle$, $|\psi_1\rangle$ with vanishing overlap. In the intermediate stage, the particles approach each other and scatter. In the final stage, long after the collision, the particles are again far apart and are detected by one-particle detectors. In the initial and final stage we perform essentially one-particle preparations/measurements, i.e. we have the choice to treat the particles as distinct, whereas in the intermediate stage the collision is correctly described only if we apply a symmetric evolution operator to a properly (anti-)symmetrized two-particle state, that is, if we treat the two particles as identical. The existence of two equivalent descriptions (``distinct" versus ``identical") of the two-particle system in the initial and final stages and the {\it correspondence rule} that relates the two is an essential ingredient of the standard quantum-mechanical formalism. In this section we want to determine how the conditions for the existence of two equivalent descriptions (``distinct" and ``identical") and the correspondence rule between the latter are modified in the twisted formalism. As a by-product, we will realize that closed systems can still be described consistently: If we e.g. want to describe a system of identical particles in our lab we are essentially allowed to forget about the existence of other particles of the same kind in the universe. Let us consider two-particle scattering again: Let initial states $|\psi_1\rangle,|\psi_2\rangle$ range on some orthogonal subspaces ${\cal H}_1, {\cal H}_2$ of the whole Hilbert space.\footnote{More generally, if the preparation were uncomplete, we would assume that particle 1,2 is in a {\it mixture} of states of ${\cal H}_1, {\cal H}_2$ respectively.} \begin{enumerate} \item[\bf(1)] We can treat the two particles as {\it distinct} particles described by the state \begin{equation} |\psi_d\rangle:= |\psi_1\rangle\otimes|\psi_2\rangle \hspace*{9mm}\in\hspace*{9mm} {\cal H}_1\otimes {\cal H}_2. \label{stated} \end{equation} A measurement process is described via a two-particle observable ${\cal O}_1\otimes {\cal O}_2$, ${\cal O}_i:{\cal H}_i\rightarrow {\cal H}_i$; the probability amplitude to find the two-particle system in a state $|\psi_d'\rangle:= |\psi_1'\rangle\otimes|\psi_2'\rangle$ is $\langle\psi_d|\psi_d'\rangle=\langle\psi_2|\psi_2'\rangle\langle\psi_1|\psi_1'\rangle$. This amounts respectively to measuring ${\cal O}_1$ on the first {\it and} ${\cal O}_2$ on the second, and to the probability amplitude to find particle 1 in state $|\psi_1'\rangle$ {\it and} particle 2 in state $|\psi_2'\rangle$. In particular, setting ${\cal O}_2=id$, $|\psi_2'\rangle=|\psi_2\rangle$ means that we neglect the information that we have on the second particle, \mbox{\it i.e.\/ } we ignore its existence. \item[\bf(2)] We can treat them as {\it identical} particles forming a two-particle system and describe the latter by the twisted (anti)symmetrized state \begin{equation} |\psi\rangle= P^{F_{12}}_{S/A}|\psi_d\rangle:=\frac {F_{12}}{\sqrt{2}}\left( |\psi_1\rangle\otimes|\psi_2\rangle \pm |\psi_2\rangle \otimes|\psi_1\rangle\right)\in ({\cal H}\otimes{\cal H})^{F_{12}}_{\pm} \label{sstate} \end{equation} of bosons $(+)$ or fermions $(-)$. The measurement process of (1) is now described by acting on $|\psi\rangle$ through the twisted symmetrized two-particle observable $F_{12}\left({\cal O}_1\otimes {\cal O}_2+{\cal O}_2\otimes {\cal O}_1\right)F_{12}^{-1}\in (\mbox{$\cal A$}\otimes A)^{F_{12}}_+$. \end{enumerate} Description (2) is perfectly equivalent to (1) because the mapping (1)$\rightarrow$(2) preserves scalar products between states (\mbox{\it i.e.\/ } probability amplitudes: $\langle\psi'|\psi\rangle=\langle\psi_d'|\psi_d\rangle= \langle\psi_1'|\psi_1\rangle\langle\psi_2'|\psi_2\rangle)$ and spectra of the observables (\mbox{\it i.e.\/ } results of measurements). For the dynamical evolution, including the collision, it is necessary to use description (2), which involves in an essential way the quantum statistics. Nevertheless, if for later times the state $|\psi\rangle(t)$ becomes a combination of states of the form \begin{equation} |\psi\rangle(t)=\sum\limits_{i,j}a_{ij} \frac {F_{12}}{\sqrt{2}}\left( |i\rangle\otimes|j\rangle\pm |j\rangle\otimes|i\rangle\right), \end{equation} where $|i\rangle\in {\cal H}_1'$, $|j\rangle\in {\cal H}_2'$, and ${\cal H}_1'$, ${\cal H}_2'$ are {\it orthogonal} subspaces of ${\cal H}$ (describing e.g. the states of the particle in detectors 1,2 respectively) then description (1) can be implemented again: we can apply $F^{-1}_{12}$ and drop the (anti-)symmetrization to get the state \begin{equation} |\psi_d\rangle(t)=\sum\limits_{i,j}a_{ij} |i\rangle\otimes|j\rangle, \end{equation} which will give the final correlation between the potential measurements in the two detectors. The case of more than two particles can be treated in analogy to the case of two particles. Now however we will want to split the collection of particles into two (or more) {\em subsystems\/} instead of into single particles. If there is negligeable overlap between subsystems we are again not forced to treat {\em all\/} particles as identical particles; we can describe particles belonging to different subsystems as distinct, but we still have to twist (anti-)symmetrize each subsystem. If we look at the dynamical evolution, then the same considerations as in the case of two particles will apply. In particular as long as the interaction (of any kind) between a subsystems and the remaining particles is negligeable then we have the choice to consider one subsystem as isolated (implying that we forget the other particles) or of treating all particles as identical. These considerations hold also when the total number of particles of one kind is very large (virtually infinite) compared to the number in one subsystem. Take this subsystem to be our laboratory and we see that as in the standard formulation, to compute any concrete prediction we can but we don't have to consider all particles of the given type present in the universe at the same time [description ``identical''], namely we may ignore the ones ``outside our laboratory'' [description ``distinct'']. In principle however we could apply the postulates of identical particles, through description ``identical'', to {\it all} particles of the same type in the universe, without finding inconsistent predictions. In other words, the twisted postulates of Quantum Mechanics for identical particles are completely general and self-consistent. \sect{Quantum Symmetries} \label{qsym} While their introduction was shown to be consistent, there was so far no need for the $F_{12\ldots n}$. Now we take the issue of quantum group symmetries into consideration. The picture we have in mind is that of a multi-particle quantum mechanical model (consisting of identical particles) on which we would like to implement generalized (symmetry) transformations through the action of a generic Hopf algebra $H$.\footnote{Later we will concentrate on the case of a twisted image of a cocommutative (quasi-) Hopf algebra; {\em e.g.\/} $U_q(g)$.} As given data we take the constituent one-particle system, governed by a $*$-algebra $\mbox{$\cal A$}$ of operators that act on a Hilbert space ${\cal H}$, a $*$-Hopf algebra $H$ with coproduct $\Delta$, counit $\varepsilon$, antipode $S$ and complex conjugation $*$, and a unitary realization $\rho$ of $H$ in $\mbox{$\cal A$}$. The key idea that leads to a construction of multi-particle systems that consistently transform under Hopf algebra actions is that properties of the coproduct should have to do with (twisted) (anti-)symmetry of states and operators. We will find that coproducts should be considered as being (twisted) symmetric---even when we are dealing with non-cocommutative Hopf algebras as symmetries. Let us start by recalling what it means that a one-particle system transforms under the action of $H$. \subsection{One-Particle Transformations} \label{onep} To begin, we need a representation $\rho$ of $H$ on ${\cal H}$ which realizes $H$ in $\mbox{$\cal A$}$:\footnote{A given algebra of operators might first have to be extended for this scope.} \begin{equation} \rho\, : \: H \: \rightarrow \: \mbox{$\cal A$}; \end{equation} the map $\rho$ is linear and an algebra homomorphism $\rho(x y) = \rho(x)\rho(y)$; $\rho(1_H) = 1_{\cal A}$ is the identity operator on ${\cal H}$. It is called a unitary representation if in addition \begin{equation} \rho(x)^* = \rho(x^*). \label{unitrep} \end{equation} (For a representation that is not unitary we would find in contrast $\rho(x)^* = \overline{\rho^\vee}( x^*)$, where $\overline{\rho^\vee}$ is the complex conjugate of the contragredient representation. For a matrix representation: $(T^\vee)^i{}_j = S(T^j{}_i) = (T^{-1})^j{}_i$. ) Let $x \in H$, $\mbox{$\cal O$} \in \mbox{$\cal A$}$ and $|\psi\rangle \in {\cal H}$. The actions of $x$ on the one-particle states $|\psi\rangle$ and and $\mbox{$\cal O$} |\psi\rangle$ are given via $\rho$ \begin{eqnarray} x \triangleright |\psi\rangle & = & \rho(x) |\psi\rangle,\label{actsta1}\\ x \triangleright \big(\mbox{$\cal O$} |\psi\rangle \big) & = & \rho(x) \mbox{$\cal O$} |\psi\rangle, \label{actsta2} \end{eqnarray} while on the other hand the action of $x$ on the product $\mbox{$\cal O$} |\psi\rangle$ (that is, on an element of the bigger $H$-module containing both $\mbox{$\cal A$}$ and ${\cal H}$) should be computed with the coproduct $\Delta$, \mbox{\it i.e.\/ } \begin{equation} x \triangleright \big(\mbox{$\cal O$} |\psi\rangle \big) = (x_{(1)} {\,\stackrel{s}{\triangleright}\,} \mbox{$\cal O$}) (x_{(2)} \triangleright |\psi\rangle ). \label{actsta3} \end{equation} Here and in the sequel we will use Sweedler's notation $\Delta(x)\equiv x_{(1)}\otimes x_{(2)} $ for the coproduct (in the RHS a sum $\sum_i x^i_{(1)}\otimes x^i_{(2)}$ of many terms is implicitly understood); similarly, $\Delta^{(n-1)}(x) \equiv x_{(1)}\otimes\ldots\otimes x_{(n)}$ for the $(n\!-\!1)$-fold coproduct in Sweedler's notation. As known, it follows that the action of $H$ on the one-particle operator $\mbox{$\cal O$}$ is given by\footnote{See however the remark on page~\ref{rem}.} \begin{equation} x {\,\stackrel{s}{\triangleright}\,} \mbox{$\cal O$} = \rho(x_{(1)}) \:\mbox{$\cal O$}\: \rho(S x_{(2)}), \qquad x \in H,\ \mbox{$\cal O$} \in \mbox{$\cal A$}. \label{adjoint} \end{equation} As a concrete example, the reader may think of the case of quantum mechanics in ordinary three-dimensional space with transformations consisting of ordinary rotations; in that case $H$ is the (undeformed) universal enveloping algebra $U(su(2))$ of the (covering of the) Lie group $SO(3)$. $\rho$ maps elements of $U (su(2))$ into operators acting on ${\cal H}$, out of which we can single out unitary operators ``$U$'' realizing finite rotations (i.e. elements of $SO(3)$), as well as hermitean ones ``$x$'' realizing infinitesimal rotations (i.e. elements of $su(2)$) and generating the whole algebra; in these two cases the action (\ref{adjoint}) reduces respectively to conjugation $U\mbox{$\cal O$} U^{-1}$ and to taking the commutator $[ix,\mbox{$\cal O$}]$. A rotation symmetry of the Hamiltonian usually turns elements of $\rho(U(su(2))$ (\mbox{\it e.g.\/ } angular momentum components) into useful observables for studying the dynamics of the system. \subsubsection{Unitary Transformations} Hermitean conjugation turns an element of ${\cal H}$, a ``ket'', into a ``bra'' which lives in ${\cal H}^*$ and transforms under the contragredient representation. This picture should be preserved under transformations. As we know, in the classical case only unitary and---in the infinitesimal case---anti-hermitean transformation operators have the required property. In the general Hopf algebra case the required property is $S(x) = x^*$; we will call such elements of $H$ {\em quantum unitary}. We stress the point that there are two notions of unitarity which should not be confused: that of a representation, and that of a transformation. Quantum unitary elements also leave the $*$-structure of $\mbox{$\cal A$}$ invariant \cite{thesis}. The condition for an element $u \in H$ to satisfy \begin{equation} (u {\,\stackrel{s}{\triangleright}\,} \mbox{$\cal O$})^* = u {\,\stackrel{s}{\triangleright}\,} \mbox{$\cal O$}^* \qquad \forall \mbox{$\cal O$} \in A \label{requ} \end{equation} is again \begin{equation} u^* = S(u) \qquad \label{quop}\mbox{(quantum unitary operator)}. \end{equation} This is seen as follows: $*$-conjugating both sides of equation (\ref{adjoint}) we find a condition \begin{equation} \rho(S u_{(2)})^* \otimes \rho(u_{(1)})^* \stackrel{!}{=} \rho(u_{(1)}) \otimes \rho(S u_{(2)}), \end{equation} or, using that $\rho$ is a unitary representation, \begin{equation} (S u_{(2)})^*\otimes (u_{(1)})^* \stackrel{!}{=} u_{(1)}\otimes S u_{(2)}. \end{equation} Taking the counit $(\varepsilon \otimes id)$ of this equation gives condition (\ref{quop}). A straightforward calculation that uses again unitarity of the representation $\rho$ and standard facts about $*$-Hopf algebras, like \ $* \circ S = S^{-1} \circ *$ shows that condition (\ref{quop}) is in fact sufficient for (\ref{requ}). {\em Remark:}\/ There exist pathological Hopf algebras (\mbox{\it e.g.\/ } with $\tau\circ\Delta = (id \otimes S^2)\Delta$) that are not $*$-Hopf algebras but still allow unitary transformations in a non-standard way. \subsection{Multi-Particle Transformations} To implement symmetry transformations (the action of $H$) on multi-particle systems one makes use of the coproduct of $H$, which enters the game in two essentially different ways. First, the coproduct is needed to extend the action of $H$ from one-particle {\it states} to $n$-particle states in a way that preserves the twisted (anti)-symmetry of identical particle states. This will constrain the choice of $F$ in section~\ref{twistmult}, and consequently also the twisted symmetry of operators, according to formula (\ref{twistop}). On the other hand, the coproduct also enters the action of $H$ on single and multiparticle operators $\mbox{$\cal O$}^{(n)}$ [see formula (\ref{adjoint}) for the one-particle case]; if the particles are identical this action should again preserve the twisted symmetry of the operators. It turns out that both consistency requirements can be simultaneously satisfied through an appropriate choice of the $F$'s. \subsubsection{Transformation of States} \label{transta} We have so far required that ${\cal H}$ be a $*$ $H$-module, \mbox{\it i.e.\/ } that it carries a $*$ representation of $H$. The main task in constructing Hilbert spaces for identical particles is then to find an operation of twist (anti-) sym\-me\-tri\-za\-tion that is compatible with the action of $H$, \mbox{\it i.e.\/ } compatible with the quantum symmetry transformations. The action of $H$ on a multi-particle Hilbert space is given once $\rho^{(n)}$ is known. A representation $\rho$ on the 1-particle Hilbert space extends to a unitary representation on the $n$-particle Hilbert space via the $(n-1)$-fold coproduct of $H$: \begin{equation} \rho^{(n)} = \rho^{\otimes n} \circ \Delta^{(n-1)}\,: \: H \: \rightarrow \: \mbox{$\cal A$}^{\otimes n} \,: {\cal H}^{\otimes n} \: \rightarrow {\cal H}^{\otimes n}. \end{equation} If $\rho$ is unitary then so is $\rho^{(n)}$, $\rho^{(n)}(x)^{*_n} = \rho^{(n)}(x^*)$, because $(* \otimes *) \circ \Delta= \Delta \circ *$. Let $x \in H$ and $|\psi^{(n)}\rangle \in {\cal H}^{\otimes n}$, then \begin{equation} x \triangleright |\psi^{(n)}\rangle = \rho^{(n)}(x)|\psi^{(n)}\rangle = \rho(x_{(1)}) \otimes\ldots\otimes \rho(x_{(n)})|\psi^{(n)}\rangle. \label{actstam} \end{equation} As always we will first consider the case of two particles. Similar considerations will apply to the case of $n\ge 3$ particles. Let $P_{12}$ be the permutation operator on ${\cal H} \otimes {\cal H}$. \paragraph{Symmetric coproduct} In the case of a {\em co-commutative}\/ (\mbox{\it i.e.\/ } symmetric under permutation) coproduct we have $$ P_{12}\: \Big((\rho \otimes \rho)\Delta_c(x)\Big) = \Big((\rho \otimes \rho)\Delta_c(x)\Big)\: P_{12} $$ and hence $$ P_{12} (x \triangleright |\psi^{(2)}\rangle) = x \triangleright (P_{12}|\psi^{(2)}\rangle). $$ This fact allows us to define symmetrizers $P_S = {1 \over 2}(I + P_{12})$ and anti-symmetrizers $P_A = {1 \over 2}(I - P_{12})$ that commute with the action of $x$, and (anti-) symmetrized Hilbert spaces \begin{eqnarray} P_S({\cal H} \otimes {\cal H}) & \equiv & ({\cal H} \otimes {\cal H})_+,\\ P_A({\cal H} \otimes {\cal H}) & \equiv & ({\cal H} \otimes {\cal H})_-, \end{eqnarray} that are invariant under the action of $x$. This happens for instance if $H = U(\mbox{\bf g})$, $\mbox{\bf g} = ${\it Lie}$(G)$. Then $U(\mbox{\bf g})$ is generated by primitive elements $X_i$ with coproduct \begin{equation} \Delta^{(n)}(X_i) = \Delta_c^{(n)}(X_i) = X_i \otimes 1 \otimes\ldots\otimes 1 + \ldots + 1 \otimes\ldots\otimes X_i; \end{equation} $\Delta_c^{(n)}(X_i)$ is invariant under permutations and we can set $F_{12\ldots n} ={\bf 1}\otimes\ldots\otimes {\bf 1}$. \paragraph{Deformed coproduct} If the coproduct is {\em not co-commutative}, as it happens for a generic Hopf algebra, then the problem arises that the action of $H$ on $({\cal H} \otimes {\cal H})$ will no more preserve the subspaces $({\cal H} \otimes {\cal H})_{\pm}$. While we should not change the form of the coproduct (it is at the very heart of quantum groups and tells us how to act on tensor products) we may however modify our notion of symmetric operators and (anti-) symmetrized Hilbert spaces. We can require that \begin{equation} \rho^{(n)}(H) \subset (\underbrace{\mbox{$\cal A$} \otimes\ldots\otimes\mbox{$\cal A$}}_{\mbox{$n$-times}})_+^{F_{12\ldots n}} \label{delop} \end{equation} for some $F_{12\ldots n}$, so that the system of $n$ identical particles carries a $*$-representation of $H$ as well. This is certainly satisfied if \begin{equation} \rho^{(n)}(X) = F_{12..n} \rho^{(n)}_c(X) F_{12..n}^{-1}, \label{deltwist} \end{equation} where $\rho^{(n)}_c:=\rho^{\otimes n} \circ \Delta^{(n-1)}$ and $\Delta_c$ is a co-commutative coproduct. Equation (\ref{deltwist}) has to be read as a condition on both $\Delta_c$ and $F_{12..n}$. If $H = \mbox{$U_{q}{\/\mbox{\bf g}}$}$ \cite{dr,ji,frt}, where $\mbox{\bf g}$ is the Lie algebra of one of the simple Lie groups of the classical series, the following theorem due to Drinfel'd and Kohno will be our guidance to the correct choice of the $F$'s we need to satisfy equations (\ref{delop}) and (\ref{deltwist}): \noindent {\bf Drinfel'd Proposition 3.16 in Ref. \cite{Drinfeld}} \begin{enumerate} {\em \item There exists an algebra isomorphism $\phi: \mbox{$U_{q}{\/\mbox{\bf g}}$} \stackrel{~}{\leftrightarrow} (U \mbox{\bf g})([[h]])$, where $h = \ln q$ is the deformation parameter. \item If we identify the isomorphic elements of $\mbox{$U_{q}{\/\mbox{\bf g}}$}$ and $(U \mbox{\bf g})([[h]])$ then there exists an $\mbox{$\cal F$} \in \mbox{$U_{q}{\/\mbox{\bf g}}$} \otimes \mbox{$U_{q}{\/\mbox{\bf g}}$}$ such that: \begin{equation} \Delta(a) = \mbox{$\cal F$} \Delta_c(a) \mbox{$\cal F$}^{-1},\hspace*{9mm} \forall a \in \mbox{$U_{q}{\/\mbox{\bf g}}$} \stackrel{~}{=} (U \mbox{\bf g})([[h]]) \label{fdf} \end{equation} where $\Delta$ is the coproduct of $\mbox{$U_{q}{\/\mbox{\bf g}}$}$ and $\Delta_c$ is the (co-commutative) coproduct of $U(\mbox{\bf g})$. \item $(U \mbox{\bf g})([[h]])$ is a quasi-triangular quasi-Hopf algebra (QTQHA) with universal $\mbox{$\cal R$}_\Phi = q^{t/2}$ and a quasi-coassociative structure given by an element $\Phi \in \left((U \mbox{\bf g})^{\otimes 3}([[h]])\right)$ that is expressible in terms of $\mbox{$\cal F$}$. $(U \mbox{\bf g})([[h]])$ as QTQHA can be transformed via the twist by $\mbox{$\cal F$}$ into the quasi-triangular Hopf algebra $\mbox{$U_{q}{\/\mbox{\bf g}}$}$; in particular, the universal $\mbox{$\cal R$}$ of $\mbox{$U_{q}{\/\mbox{\bf g}}$}$ is given by $\mbox{$\cal R$}= \mbox{$\cal F$}_{21}\mbox{$\cal R$}_\Phi \mbox{$\cal F$}^{-1}$. } \end{enumerate} Here $(U \mbox{\bf g})([[h]])$ denotes the algebra of formal power series in the elements of a basis of $\mbox{\bf g}$, with coefficients being entire functions of $h$; $(U \mbox{\bf g})([[h]])|_{h=0}=U\mbox{\bf g}$. Point 1) essentially says that it is possible to find $h$-dependent functions of the generators of $U\mbox{\bf g}$ which satisfy the algebra relations of the Drinfel'd-Jimbo generators of $\mbox{$U_{q}{\/\mbox{\bf g}}$}$ and vice versa. We recall here that the quasi-triangular Hopf algebras $\mbox{$U_{q}{\/\mbox{\bf g}}$}$ can be obtained as quantizations of Poisson-Lie groups associated with solutions of the modified classical Yang-Baxter equations (MCYBE) corresponding to $\mbox{\bf g}$. If the Hopf algebra $H$ can be obtained as the quantization of a Poisson-Lie group associated with a solution of the classical Yang-Baxter equation (CYBE) corresponding to some $\mbox{\bf g}$,\footnote{In this case $H$ is is triangular, \mbox{\it i.e.\/ } $\mbox{$\cal R$}_{21}\mbox{$\cal R$}_{12}={\bf 1}$} then another (and chronologically preceding) theorem by Drinfel'd \cite{drinf} states the existence of a different $\mbox{$\cal F$}$ with similar properties as in the previous theorem---except that now it is enough to twist $(U \mbox{\bf g})([[h]])$ equipped with the ordinary {\it coassociative} structure in order to obtain $H$. The quasi-coassociative structure $\Phi$ and the quasi-triangular structure $\mbox{$\cal R$}_\Phi$ of point 3) in the theorem reduce to $\Phi={\bf 1}\otimes {\bf 1} \otimes {\bf 1}$, $\mbox{$\cal R$}_\Phi={\bf 1}\otimes {\bf 1}$ ; the universal $\mbox{$\cal R$}$ is given by $\mbox{$\cal R$}= \mbox{$\cal F$}_{21} \mbox{$\cal F$}^{-1}$. A simple introduction to these topics can be found for instance in Ref. \cite{Taktadjan}. As shown in Ref. \cite{jurco}, one can always choose a unitary $\mbox{$\cal F$}$, if $H$ is a compact section of $U_q(g)$ (\mbox{\it i.e.\/ } when $q\in {\bf R}$). If the $\mbox{$\cal F$}$ one starts with is not unitary, when one simply multiplies it with the invariant (!) tensor $(\mbox{$\cal F$}^* \mbox{$\cal F$})^{-1/2}$ to obtain a new $\tilde{\mbox{$\cal F$}}$ that is unitary. These theorems suggest that one can use the unitary twisting operator $\mbox{$\cal F$}$ to build $F_{12}$ for a 2-particle sytem. For example: \begin{enumerate} \item If $\mbox{$\cal A$} =\rho( \mbox{$U_{q}{\/\mbox{\bf g}}$})$, then we choose $$ F =\rho^{\otimes 2}( \mbox{$\cal F$}) .$$ \item If $\mbox{$\cal A$} =$ classical Heisenberg algebra $\otimes U^{\mbox{spin}}(su(2)) \otimes \rho(\mbox{$U_{q}{\/\mbox{\bf g}}$})$, were $\mbox{$U_{q}{\/\mbox{\bf g}}$}$ plays the role of an internal symmetry, then we can set $$ F_{12} = \mbox{id}^{(2)}_{\mbox{Heisenberg}} \otimes \mbox{id}^{(2)}_{\mbox{spin}} \otimes \rho^{\otimes 2}( \mbox{$\cal F$}) $$ \item If $\mbox{$\cal A$}$ is the q-deformed Poincar\'e algebra of ref. \cite{wess2,majid}, and $H$ is the corresponding q-deformed Lorentz Hopf algebra, realized through $\rho$ in $\mbox{$\cal A$}$, then we can again define $$ F_{12} = \rho^{\otimes 2}( \mbox{$\cal F$}),$$ where $\mbox{$\cal F$}$ belongs to the homogeneous part. The same applies for other inhomogeneous algebras, like the q-Euclidean ones, constructed from the braided semi-direct product \cite{majid} of a quantum space and of the corresponding homogeneous quantum group. For both of these examples the one-particle representation theory is known \cite{wess2,fioeu}. \end{enumerate} For $n$-particle systems one can set $F_{12...n}=\rho^{\otimes n}( \mbox{$\cal F$}_{12...n})$, where now we should choose one particular element $\mbox{$\cal F$}_{12...n}$ of $H^{\otimes n}$ satisfying the condition \begin{equation} \Delta(x)=\mbox{$\cal F$}_{12...n}\Delta_c(x)(\mbox{$\cal F$}_{12...n})^{-1}. \label{condition} \end{equation} To obtain one such $\mbox{$\cal F$}_{12...n}$ it is enough to act on eq. (\ref{fdf}) $(n-2)$ times with the coproduct in some arbitrary order. When $n=3$, for instance, one can use either $\mbox{$\cal F$}'_{123}:=[(\Delta\otimes id)(\mbox{$\cal F$})]\mbox{$\cal F$}_{12}$ or $\mbox{$\cal F$}''_{123}:=[(id\otimes\Delta)(\mbox{$\cal F$})]\mbox{$\cal F$}_{23}$. These two elements coincide in the case previously mentioned of Hopf algebras associated to solutions of the CYBE, as proved by Drinfeld \cite{drinf}. In the the case of $\mbox{$U_{q}{\/\mbox{\bf g}}$}$, they do not coincide, but nevertheless $\Phi:=\mbox{$\cal F$}''_{123}(\mbox{$\cal F$}'_{123})^{-1}\neq {\bf 1}\otimes{\bf 1}\otimes{\bf 1}$ commutes with $\Delta^{(2)}(H)$. In section (\ref{example}) we will show (in the $U_q(su(2))$ case) how to find a continuous family of $\mbox{$\cal F$}_{123}$ interpolating between $\mbox{$\cal F$}'_{123}$ and $\mbox{$\cal F$}''_{123}$. \paragraph{Note:} {}From (\ref{fdf}) follows $(\tau\circ\Delta)(a) = {\cal M} \Delta(a) {\cal M}^{-1}$ with ${\cal M} := \mbox{$\cal F$}_{21} \mbox{$\cal F$}^{-1}.$ This is not the usual relation $(\tau\circ\Delta)(a) = \mbox{$\cal R$} \Delta(a) \mbox{$\cal R$}^{-1}$ of a quasi-triangular Hopf algebra; the latter is rather obtained by rewriting equation (\ref{fdf}) in the form $\Delta(a) = \mbox{$\cal F$} q^{t/2} \Delta_c(a) q^{-t/2} \mbox{$\cal F$}^{-1}$ where $t = \Delta_c(C_c) - 1\otimes C_c - C_c \otimes 1$ is the invariant tensor \ ($[t , \Delta_c(a) ] = 0 \hspace{3mm}\forall\hspace{3mm} a \in U \mbox{\bf g}$) \ corresponding to the Killing metric, and $C_c$ is the quadratic casimir of $U\mbox{\bf g}$. ${\cal M}$, unlike $\mbox{$\cal R$}$, has not nice properties under the coproducts $\Delta\otimes id$, $id\otimes \Delta$. The reader might wonder whether we could use equation $[P_{12}R, (\mbox{$\cal A$}\otimes\mbox{$\cal A$})'_+]=0$ (where $R=\rho^{\otimes 2}(\mbox{$\cal R$})$), instead of eq. (\ref{commop}), to single out a modified symmetric algebra $(\mbox{$\cal A$}\otimes\mbox{$\cal A$})'_+ \subset \mbox{$\cal A$}\otimes\mbox{$\cal A$}$; in fact, the former is also an equation fulfilled by $\rho^{\otimes 2}(\Delta(H))$ and reduces to the classical eq. (\ref{symop}) in the limit $q\rightarrow 1$. However $[P_{12}R, (\mbox{$\cal A$}\otimes\mbox{$\cal A$})'_+]=0$ is fulfilled {\it only} by the sub-algebra $\rho^{\otimes 2}(\Delta(H))\subset(\mbox{$\cal A$}\otimes\mbox{$\cal A$})$ itself, essentially because $q^{t / 2}$ does not commute with {\em all}\/ symmetric operators, but only with the ones corresponding to coproducts. Therefore, $(\mbox{$\cal A$}\otimes\mbox{$\cal A$})_+'$ defined via $P_{12}R$ (instead of $P_{12}^{F_{12}}$) is not big enough to be in one-to-one correspondence with the classical $(\mbox{$\cal A$}\otimes\mbox{$\cal A$})_+$, \mbox{\it i.e.\/ } is not suitable for our purposes. Explicit universal $\mbox{$\cal F$}$'s for $\mbox{$U_{q}{\/\mbox{\bf g}}$}$ are not given in the literature, up to our knowledge; an explicit universal $\mbox{$\cal F$}$ for a family of deformations (which include quantizations of solutions of both of a CYBE and of a MCBYE) of the Heisenberg group in one dimension was given in Ref. \cite{bonechi}. However, for most practical purposes one has to deal with representations $F$ of $\mbox{$\cal F$}$. A general method for constructing the matrices $F$ acting on tensor products of two arbitrary irreducible representations of compact sections of $\mbox{$U_{q}{\/\mbox{\bf g}}$}$ is presented in Ref. \cite{eng}---there explicit formulas are given for the $A,B,C,D$-series in the fundamental representation. In \cite{cuza} matrices twisting the classical coproduct into the $q$-deformed one were constructed from $q$-Clebsch-Gordan coefficients. Moreover, in the intrinsic formulation of the twisted (anti-)symmetrization postulates [eqs. (\ref{astate}) -- (\ref{asop})] one only needs the twisted permutators $P_{12\ldots n}^{F_{12\ldots n}}$ (not the $F_{12\ldots n}$ themselves); explicit universal expressions for the latter can be found much more easily, as we show in section \ref{example} for $P_{12}^{\tiny\mbox{$\cal F$}_{12}}$ in the case $H=U_q(su(2))$. We conclude that the quantum symmetry is compatible with identical particle {\em states}\/ in the twisted multi-particle description. \subsubsection{Transformation of Operators} \label{operators} Now we want to see if a consistent transformation of the twisted-symmetric operators can be defined. As we have seen in section \ref{onep}, the action on one-particle operators which makes eq. (\ref{actsta3}) consistent with eq. (\ref{actsta2}) looks formally like the quantum adjoint action. A subtle but important change in the definition of the action on multi-particle operators is needed in order to reach the same goal for multi-particle systems. Our task in this section is twofold: first we have to find the right action of the Hopf algebra $H$ for tensor products of $\mbox{$\cal A$}$, then we have to show that the definition of ``twist symmetric'' operators (associated to identical particles) is stable under this action. As before, we assume that $\rho$ is a unitary representation that realizes the Hopf algebra $H$ of transformations in $\mbox{$\cal A$}$. Let $\mbox{$\cal O$}^{(n)} \in \mbox{$\cal A$}^{\otimes n}$ (or a properly symmetrized subspace), $|\psi_n\rangle\in{\cal H}^{\otimes n}$ (or a properly (anti)symmetrized subspace); we require, as in the one-particle case, \begin{equation} (x_{(1)} {\,\stackrel{s}{\triangleright}\,} \mbox{$\cal O$}^{(n)}) (x_{(2)} \triangleright |\psi_n\rangle ) = x \triangleright \big(\mbox{$\cal O$}^{(n)} |\psi_n\rangle \big) = \rho^{(n)}(x) \mbox{$\cal O$}^{(n)} |\psi_n\rangle. \label{actsta4} \end{equation} Recalling eq. (\ref{actstam}) it is easy to see that to satisfy this goal the action (\ref{adjoint}) has to generalize to multi-particle operators in the following way: \begin{equation} \begin{array}{rcl} x {\,\stackrel{s}{\triangleright}\,} \mbox{$\cal O$}^{(n)} & = & \rho^{(n)}(x_{(1)})\: \mbox{$\cal O$}^{(n)} \:\rho^{(n)}(S x_{(2)})\\ & = & \rho^{\otimes n} \big(x_{(1)}\otimes \ldots \otimes x_{(n)}\big) \:\mbox{$\cal O$}^{(n)} \:\rho^{\otimes n}\big(S x_{(2n)}\otimes \ldots \otimes S x_{(n+1)}\big).\label{symmetry} \end{array} \end{equation} {\em Remark:}\ In the case that $\mbox{$\cal O$} = \rho(y)$ with $y \in H$ the action on one-particle operators is nothing but the adjoint action $x \stackrel{\mbox{\scriptsize ad}}{\triangleright} y = x_{(1)} y S(x_{(2)})$. The action on multi-particle operators is however different: For instance in the case that $ \mbox{$\cal O$}^{(2)} = (\rho\otimes\rho)(y_i\otimes y^i)$ with $y_i\otimes y^i \in H \otimes H$ we get $$x {\,\stackrel{s}{\triangleright}\,} (y_i\otimes y^i) = x_{(1)} y_i S x_{(4)} \otimes x_{(2)} y^i S x_{(3)}$$ and {\em not}\/ $$ x \stackrel{\mbox{\scriptsize ad}}{\triangleright} (y_i\otimes y^i) = x_{(1)} \stackrel{\mbox{\scriptsize ad}}{\triangleright} y_i \otimes x_{(2)}\stackrel{\mbox{\scriptsize ad}}{\triangleright} y^i = x_{(1)} y_i S x_{(2)} \otimes x_{(3)} y^i S x_{(4)} $$ as one might have expected. Both actions ``$\stackrel{\mbox{\scriptsize ad}}{\triangleright}$'' and ``${\,\stackrel{s}{\triangleright}\,}$'' coincide for co-commutative coproducts. The former action treats multi-particle operators as tensor products of $H$-modules, the latter action is related to the natural Hopf algebra structure on $\Delta(H)$ that is given in Sweedler's book \cite{Sweedler}. Briefly, Sweedler's argument is the following. For any given number $n$, $\Delta^{(n-1)}(H)$ can be viewed as a Hopf algebra with a natural coproduct. Now formula (\ref{adjoint}) is applicable for any $n$---we just have to take care to use the natural Hopf algebra structure for each of the $\Delta^{(n-1)}(H)$.\footnote{The action ``${\,\stackrel{s}{\triangleright}\,}$'' was also used in Ref.~\cite{SWZ} to define covariance properties of tensors in $H^{\otimes n}$} The notion of unitary multi-particle transformations generalizes to $n$ particles in an obvious way, \begin{equation} (u {\,\stackrel{s}{\triangleright}\,} \mbox{$\cal O$}^{(n)})^* = u {\,\stackrel{s}{\triangleright}\,} (\mbox{$\cal O$}^{(n)})^* \qquad \forall \mbox{$\cal O$}^{(n)} \in A \end{equation} and again is satisfied if $u^* = S(u)$. We now want to show that the transformation we have found is compatible with the symmetrization of operators in the twisted multi-particle description. \paragraph{Symmetric coproduct} First consider the co-commutative case. Let $$ (\mbox{$\cal A$}\otimes\ldots\otimes\mbox{$\cal A$})_+ = \{ a \in \mbox{$\cal A$}\otimes\ldots\otimes\mbox{$\cal A$} : [P_{i,i+1} , a] =0 , i=1,\ldots n-1\} $$ be the completely symmetrized space of $n$-particle operators. In the case of a co-commutative \mbox{\it i.e.\/ } {\em symmetric}\/ coproduct $\Delta_c$ any of the permutation operators $P_{i,i+1}$ will commute with the action (\ref{symmetry}): \begin{equation} \begin{array}{rcl} \Big[P_{i,i+1} \, , \, \big(x {\,\stackrel{s}{\triangleright}\,} \mbox{$\cal O$}^{(n)}\big) \Big] & = & \Big[ P_{i,i+1} \, , \,\rho^{\otimes n} \Big( \Delta_c^{(n-1)}(x_{c(1)}) \Big) \:\mbox{$\cal O$}^{(n)}\: \rho^{\otimes n}\Big(\Delta_c^{(n-1)}(S_c x_{c(2)})\Big)\Big]\\ & = & x {\,\stackrel{s}{\triangleright}\,} \big[P_{i,i+1} \, , \,\mbox{$\cal O$}^{(n)}\big], \qquad \mbox{for } \Delta_c \mbox{ cocommutative }. \end{array} \label{cococ} \end{equation} Here $x_{c(1)}\otimes x_{c(2)}\equiv\Delta_c(x)$ and $S_c$ is the cocommutative antipode. \paragraph{Deformed coproduct} Let ${\mbox{$\cal F$}_{12\ldots n}}\in H^{\otimes n}$ be as in equation (\ref{condition}) namely such that $\Delta^{(n-1)}(x) = {\mbox{$\cal F$}_{12\ldots n}}\Delta_c^{(n-1)}(x) {\mbox{$\cal F$}_{12\ldots n}}^{-1}$ for all $x \in H$. As in the previous section we will use its representation $F_{12\ldots n} \equiv \rho^{\otimes n}({F_{12\ldots n}})$ for the similarity transformation of section~\ref{twistmult}. We note that relation (\ref{cococ}) also holds with $x_{c(1)}\otimes x_{c(2)}$ and $S_c$ replaced by the non-cocomutative $x_{(1)}\otimes x_{(2)}\equiv\Delta(x)$ and the corresponding antipode $S$: \begin{equation} \begin{array}{rcl} & & \Big[ P_{i,i+1} \, , \,\rho^{\otimes n} \big( \Delta_c^{(n-1)}(x_{(1)}) \big) \:\mbox{$\cal O$}^{(n)}\: \rho^{\otimes n}\big(\Delta_c^{(n-1)}(S x_{(2)})\big)\Big]\\ & & = \rho^{\otimes n} \big(\Delta_c^{(n-1)}(x_{(1)}) \big) \big[P_{i,i+1} \, , \,\mbox{$\cal O$}^{(n)}\big] \rho^{\otimes n}\big(\Delta_c^{(n-1)}(S x_{(2)})\big). \end{array} \label{mcococ} \end{equation} Conjugating this relation by $F_{12\ldots n}$ we easily find the non-cocommutative analog of equation (\ref{cococ}), because $\rho^{(n)}(x) = \rho^{\otimes n}(\Delta^{(n-1)}(x)) = F_{12\ldots n} \rho^{\otimes n}(\Delta_c^{(n-1)}(x)) F^{-1}_{12\ldots n}$: \begin{equation} \Big[P^{F_{12\ldots n}}_{i,i+1} \, , \, \big(x {\,\stackrel{s}{\triangleright}\,} \mbox{$\cal O$}^{(n)}\big) \Big] = x {\,\stackrel{s}{\triangleright}\,} \big[P^{F_{12\ldots n}}_{i,i+1} \, , \,\mbox{$\cal O$}^{(n)}\big] \qquad \forall x \in H. \label{noncococ} \end{equation} Consequently, since the LHS vanishes if the RHS does: \begin{equation} H\,:\: (\mbox{$\cal A$}\otimes\ldots\otimes\mbox{$\cal A$})_+^{F_{12\ldots n}}\:\rightarrow\: (\mbox{$\cal A$}\otimes\ldots\otimes\mbox{$\cal A$})_+^{F_{12\ldots n}}. \end{equation} The quantum symmetry is hence compatible with identical particle {\em operators}\/ in the twisted multi-particle description.\\[1em] {\em Remark:}\ The transformation (\ref{symmetry}) is not the only one \label{rem} compatible with the twisted symmetrization. The important point is that the transformation must be based on $\rho^{(n)}(x) = \rho^{\otimes n}(\Delta^{(n-1)}(x))$. The ordinary commutator $\big[\rho^{(n)}(x)\, , \, \mbox{$\cal O$}^{(n)}\big]$ also leaves $(\mbox{$\cal A$}^{\otimes n})^{F_{12\ldots n}}_+$ invariant, simply because $\rho^{(n)}(x) \in (\mbox{$\cal A$}^{\otimes n})^{F_{12\ldots n}}_+$. These two transformations usually coincide in ordinary quantum mechanics. Here they have different interpretations: Let $h \subset H$ be a sub-algebra of $H$. The operator $\mbox{$\cal O$}^{(n)}$, $n\ge 1$, is symmetric (\mbox{\it i.e.\/ } invariant) under the transformations generated by $x\in h$ if \begin{equation} x {\,\stackrel{s}{\triangleright}\,} \mbox{$\cal O$}^{(n)} = \mbox{$\cal O$}^{(n)} \epsilon(x); \label{invar} \end{equation} it may be simultaneously diagonalizable with elements in $h$ if \begin{equation} \big[\rho^{(n)}(x)\, , \, \mbox{$\cal O$}^{(n)}\big] = 0. \label{commu} \end{equation} The two properties coincide if $\Delta(h) \subset h \otimes H$. This can be seen as follows: \begin{eqnarray} \rho^{(n)}(x) \mbox{$\cal O$}^{(n)}|\psi_n\rangle & \stackrel{(\ref{actstam})}{=}& x{\,\stackrel{s}{\triangleright}\,} (\mbox{$\cal O$}^{(n)}|\psi_n\rangle) \nonumber\\ \stackrel{(\ref{actsta4})}{=} (x_{(1)}{\,\stackrel{s}{\triangleright}\,} \mbox{$\cal O$}^{(n)})(x_{(2)}{\,\stackrel{s}{\triangleright}\,}|\psi_n\rangle) & \stackrel{(\ref{invar})}{=} &\varepsilon(x_{(1)}) \mbox{$\cal O$}^{(n)}(x_{(2)}{\,\stackrel{s}{\triangleright}\,}|\psi_n\rangle) \\ = \mbox{$\cal O$}^{(n)} (x{\,\stackrel{s}{\triangleright}\,}|\psi_n\rangle) & = & \mbox{$\cal O$}^{(n)} \rho^{(n)}(x)|\psi_n\rangle \nonumber \end{eqnarray} for any $|\psi_n\rangle\in {\cal H}^{\otimes n}$, so that eq. (\ref{invar}) implies eq (\ref{commu}); in the same way one proves the converse. The physical relevance of this case is self-evident: if both $\mbox{$\cal O$}^{(n)} $ and $\rho^{(\otimes n)}(x)$ are hermitean, then they can be diagonalized simultaneously; if one of the two, say $\rho^{(\otimes n)}(x)$, is not hermitean, given an eigenvector $|\psi_n\rangle$ of $\mbox{$\cal O$}^{(n)}$, $\rho^{(\otimes n)}(x)|\psi_n\rangle$ will be another belonging to the same eigenvalue. \sect{Explicit Example: $H=U_q(su(2))$} \label{example} We consider as a simple example of a one-particle quantum mechanical system transforming under a quantum group action the case of a q-deformed rotator, $\mbox{$\cal A$}\equiv \rho(H):=\rho[U_q(su(2))]$, with $q\in {\bf R}^+$. We determine the twisted symmetry of the systems consisting of $n\ge 2$ particles of the same kind. \subsection{$n=2$ particles} \label{example1} Let us first assume that the states of the system belong to an irreducible $*$-repre\-sent\-ation of $H$, namely ${\cal H}\equiv V_j$, where $V_j$ denotes the highest weight representation of $U_q(su(2))$ with highest weight $j=0,\frac 12,1, ...$. It is very instructive to find out what $({\cal H}\otimes {\cal H})_{\pm}^{F_{12}}$ and $(\mbox{$\cal A$}\otimes \mbox{$\cal A$})_+^{F_{12}}$ in this example are. According to point 1. of the Drinfel'd theorem, we can identify $U_q(su(2))$ and $U(su(2))$ as algebras; therefore, $V_j$ can be thought as the representation space of either one. Similarly, $V_j\otimes V_j$ can be considered as the carrier space of a (reducible) representation space of either $U_q(su(2))\otimes U_q(su(2))$ or $U(su(2))\otimes U(su(2))$; moreover, $F_{12}(V_j\otimes V_j)=V_j\otimes V_j$. Thus, we can decompose it into irreducible components either of $U_q(su(2))$ or $U(su(2))$, the operators on it being defined as $\rho^{(2)}(X)=\rho^{\otimes 2}[\Delta(X)]$ or $\rho^{(2)}_c(X)=\rho^{\otimes 2}[\Delta_c(X)]$ respectively: \begin{equation} V_j\otimes V_j=\cases{\bigoplus\limits_{0\le l\le j}\mbox{$\cal V$}_{2(j-l)}^q\oplus \bigoplus\limits_{0\le l\le j-\frac 12}\mbox{$\cal V$}_{2(j-l)-1}^q \cr \bigoplus\limits_{0\le l\le j}\mbox{$\cal V$}_{2(j-l)}\oplus \bigoplus\limits_{0\le l\le j-\frac 12}\mbox{$\cal V$}_{2(j-l)-1}; \cr} \end{equation} here $\mbox{$\cal V$}^q_J$ (resp. $\mbox{$\cal V$}_J$) denotes the irreducible component of $U_q(su(2))$ (resp. $U(su(2))$) with highest weight $J$. Moreover, from point 2) of the theorem it follows \begin{equation} F_{12}\mbox{$\cal V$}_J=\mbox{$\cal V$}_J^q, \label{vtwist} \end{equation} Let us recall now that the $\mbox{$\cal V$}_J$'s have well-defined symmetry w.r.t the permutation, namely $\mbox{$\cal V$}_{2j}, \mbox{$\cal V$}_{2(j-1)}, \ldots$ are symmetric, $\mbox{$\cal V$}_{2j-1},\mbox{$\cal V$}_{2j-3},...$ are antisymmetric. This follows from the fact that $\rho^{(2)}_c(X)$ and $P_{12}$ commute. Hence \begin{eqnarray} & & (V_j\otimes V_j)_+=\bigoplus\limits_{0\le l \le j}\mbox{$\cal V$}_{2(j-l)}\\ \nonumber & & (V_j\otimes V_j)_-= \bigoplus\limits_{0\le l\le j-\frac 12}\mbox{$\cal V$}_{2(j-l)-1}. \label{pmdecom} \end{eqnarray} {}From eq.'s (\ref{vtwist}), (\ref{pmdecom}) we finally find \begin{eqnarray} & & (V_j\otimes V_j)^{F_{12}}_+:= F_{12}(V_j\otimes V_j)_+= \bigoplus\limits_{0\le l \le j}\mbox{$\cal V$}_{2(j-l)}^q\\ \nonumber & & (V_j\otimes V_j)^{F_{12}}_-:= F_{12}(V_j\otimes V_j)_-= \bigoplus\limits_{0\le l\le j-\frac 12}\mbox{$\cal V$}_{2(j-l)-1}^q. \label{pmqdecom} \end{eqnarray} This equation says that the subspaces $\mbox{$\cal V$}^q_J\subset V_j\otimes V_j$ have well-defined ``twisted symmetry''. We can use it to build $(V_j\otimes V_j)^{F_{12}}_{\pm}$ recalling how the representations $\mbox{$\cal V$}^q_J$ are obtained. For this scope, we just have to recall the explicit algebra relations and coproduct of the generators $h,X^{\pm}$ of $U_q(su(2))$: \begin{eqnarray} [h,X^{\pm}]& = & \pm 2 X^{\pm} \qquad \qquad \qquad \qquad [X^+,X^-] = \frac{q^{h}-q^{-h}}{q- q^{-1}} \nonumber\\ \Delta(h) & = & {\bf 1}\otimes h + h\otimes {\bf 1} \qquad \qquad \Delta(X^{\pm}) = X^{\pm}\otimes q^{-\frac{h}2} + q^{\frac{h}2} \otimes X^{\pm}. \label{uqsu2comrel} \end{eqnarray} Let $\{|j,m\rangle\}_{m=-j,1-j,...j}$ be an orthonormal basis of $V_j$ consisting of eigenvectors of $\rho(\frac h2)$ with eigenvalues $m$. The generators $X^{\pm}$ can be represented in terms of this basis in the following way \begin{eqnarray} \rho(X^+)|j,m\rangle&=&\sqrt{[j+m+1]_q[j-m]_q}|j,m+1\rangle,\\ \rho(X^-)|j,m\rangle&=&\sqrt{[j-m+1]_q[j+m]_q}|j,m-1\rangle, \end{eqnarray} where $[x]_q:=\frac{q^x-q^{-x}}{q-q^{-1}}$. As well known, the highest weight vector $\Vert J,J\rangle\in \mbox{$\cal V$}^q_J$---from which the whole representation $\mbox{$\cal V$}^q_J$ can be generated by repeated applications of $\rho^{(2)}(X^-)$---is obtained by solving the equation $\rho^{(2)}(X^+)\Vert J,J\rangle=0$ for the coefficients $a_h$ of the general ansatz \begin{equation} \Vert J,J\rangle=\sum\limits_{h=\mbox{\scriptsize max}\{-j,J-j\}}^{\mbox{\scriptsize min}\{j,J+j\}} a_h |j,h\rangle\otimes |j,J-h\rangle. \end{equation} Now we are ready to understand the difference between $(H\otimes H)_+^{F_{12}}$ and its sub-algebra $\rho^{(2)}(H)$: \begin{eqnarray} \rho^{(2)}(H)&\ni& a:\mbox{$\cal V$}^q_J\rightarrow \mbox{$\cal V$}^q_J,\\ (H\otimes H)_+^{F_{12}}&\ni& b:(V_j\otimes V_j)^{F_{12}}_{\pm} \rightarrow (V_j\otimes V_j)^{F_{12}}_{\pm}. \end{eqnarray} The elements of $[\rho(H)\otimes \rho(H)]_+ \setminus \rho^{(2)}(H)$ will in general map $\mbox{$\cal V$}^q_J$ out of itself, into some $\mbox{$\cal V$}^q_{J'}$ with $J'\neq J$. If ${\cal H}$ carries a reducible $*$-representation of $H$, it will be possible to decompose it into irreducible representations $V_j$, \begin{equation} {\cal H}=\bigoplus\limits_{j\in {\cal J}}V_j \qquad\qquad {\cal J}\subset {\bf N}_0/2:=\{0,\frac 12,1\ldots\}; \end{equation} then \begin{equation} {\cal H}\otimes {\cal H}=\bigoplus\limits_{j_1,j_2\in {\cal J}}V_{j_1}\otimes V_{j _2}, \label{decom} \end{equation} and each $V_{j_1}\otimes V_{j _2}$ itself will be a representation. If $j_1=j_2$, the considerations above apply. If $j_1\neq j_2$, the irreducible components $\mbox{$\cal V$}^q_J$ ($J=|j_1-j_2|,|j_1-j_2|+1,\dots, j_1+j_2$) contained in $V_{j_1}\otimes V_{j _2}$ of course {\it will not} have well-defined symmetry (neither classical nor twisted) under permutations. However, the irreducible components $\tilde{\mbox{$\cal V$}}^q_J$ contained in $V_{j_2}\otimes V_{j _1}$ will be characterized by the same set of highest weights $J$. One can split $\mbox{$\cal V$}^q_J\oplus \tilde{\mbox{$\cal V$}}^q_J$, and therefore $V_{j_1}\otimes V_{j_2}\oplus V_{j_2}\otimes V_{j_1}$, into the direct sum of one (twisted) symmetric and one (twisted) antisymmetric components \begin{equation} \left[V_{j_1}\otimes V_{j_2}\oplus V_{j_2}\otimes V_{j_1}\right]^ {F_{12}}_{\pm}= F_{12}\frac 12[{\bf 1} \pm P_{12}]\left[V_{j_1}\otimes V_{j_2}\oplus V_{j_2}\otimes V_{j_1}\right]_{\pm} \end{equation} (the symbol $F_{12}$ has to be dropped in the untwisted case). Let $\{\Vert J,M\rangle^q_{12}\}_{M=-J,...,J}$ be an orthonormal basis of $\mbox{$\cal V$}^q_J$ consisting of eigenvectors of $\rho^{(2)}(h)$ and of $\rho^{(2)}(C_q)$ ($C_q$ denotes the casimir), and let \begin{equation} \Vert J,M\rangle^q_{12}:=\sum\limits_{m_1,m_2}{\cal C}^{j_1,j_2}_ {m_1,m_2} (J,M,q)|j_1,m_1\rangle|j_2,m_2\rangle \end{equation} be the explicit decomposition of $\Vert J,M\rangle^q_{12}$ in the tensor product basis of $V_{j_1}\otimes V_{j _2}$. Then the set $\{\Vert J,M\rangle^q_{21}\}_{M=-J,...,J}$ with \begin{equation} \Vert J,M\rangle^q_{21}:=\sum\limits_{m_1,m_2}{\cal C}^{j_2,j_1}_ {m_2,m_1} (J,M,q)|j_2,m_2\rangle|j_1,m_1\rangle, \label{basis} \end{equation} will be an orthonormal basis of $\tilde{\mbox{$\cal V$}}^q_J$ consisting of eigenvectors of $\rho^{(2)}(h)$ and of the casimir $\rho^{(2)}(C_q)$ with the same eigenvalues. Defining \begin{equation} \Vert J,M\rangle^q_{\pm}:=N\left( \Vert J,M\rangle^q_{12}\pm \Vert J,M\rangle^q_{21}\right), \qquad\qquad\qquad N^{-1}:=\sqrt{2} \end{equation} we can easily realize that $\{\Vert J,M\rangle^q_{\pm}\}_{J,M}$ is an orthonormal basis of $(V_{j_1}\otimes V_{j_2}\oplus V_{j_2}\otimes V_{j_1})^{F_{12}}_{\pm}$. Note that, if $j_1=j_2\equiv j$ and we set $N^{-1}=2$ in formula (\ref{basis}), then the vectors $\Vert J,M\rangle^q_+$ will make up the sam orthonormal basis of $V_j\otimes V_j$ as befor (they will have twisted symmetry $(-1)^{J-2j}$, see the previous case) whereas the vectors $\Vert J,M\rangle^q_-$ will vanish. We are now ready to find, as announced in sections \ref{twistmult}, \ref{qsym}, the ``universal twisted permutator'' $P_{12}^{\scriptscriptstyle {\cal F}_{12}}$ of $U_q(su(2))$, defined throug the property that the twisted permutation operator $P_{12}^{F_{12}}$ on any tensor product $V\otimes V$ [$V$ being the carrier space of a representation $\rho$ whatever of $U_q(su(2))$] can be obtained by $P_{12}^{F_{12}}=\rho^{\otimes 2} (P_{12}^{\scriptscriptstyle {\cal F}_{12}})$. We decompose $V\otimes V$ as in formula (\ref{decom}). The casimir of $U_q(su(2))$ \begin{equation} C_q=X^-X^++ \left(\frac{q^{\frac{h+1}2}-q^{\frac{-h-1}2}} {q-q^{-1}}\right)^{2} \end{equation} has eigenvalues $([j+\frac 12]_q)^2$; in the limit $q\rightarrow 1$: $C_q\rightarrow C_c+\frac14$, where $C_c$ is the usual casimir of $U(su(2))$ with eigenvalues $j(j+1)$. Defining $f(z)$ by \begin{equation} \log_q[f(z)]:=\left\{\frac 1{\ln(q)}\sinh^{-1}\left[\frac{(q-q^{-1}) \sqrt{z}}2\right]\right\}^2-\frac 14, \end{equation} it is easy to verify that $f(C_q)$ has eigenvalues $q^{j(j+1)}$. Let $\hat R:=P_{12}[\rho^{\otimes 2} (\mbox{$\cal R$})]$. Recalling the formula $\mbox{$\cal R$}={\cal F}_{21}q^{\frac t2}{\cal F}_{12}^{-1}$, we realize that the vectors $\Vert J,M\rangle^q_{\pm}\in (V_{j_1}\otimes V_{j_2}\oplus V_{j_2}\otimes V_{j_1})^{F_{12}}_{\pm}$ ($j_1\neq j_2$) are eigenvectors of $\rho^{\otimes 2}\left[f({\bf 1}\otimes C_q)f(C_q\otimes {\bf 1})\left[ f(\Delta(C_q))\right]^{-1}\right]\hat R$ and $P_{12}^{F_{12}}$ with the same eigenvalue $\pm 1$. If $j_1=j_2=j$, the same holds for the vectors $\Vert J,M\rangle^q_+$ (which form a basis of $V_j\otimes V_j$). Since this holds for all $j_1,j_2$ appearing in the decomposition (\ref{decom}), and if we let $j_1,j_2$ range on ${\cal J}$ the above vectors make up a basis of $V\otimes V$, then \begin{equation} P_{12}^{F_{12}}=f({\bf 1}\otimes\rho(C_q))f(\rho(C_q)\otimes {\bf 1}) \left[f\left(\rho^{(2)}(C_q)\right)\right]^{-1}\hat R \end{equation} on $V\otimes V$. We prefer to rewrite $\hat R$ as $\hat R=[\rho^{\otimes 2} (\mbox{$\cal R$}_{21})]P_{12}$, where $\mbox{$\cal R$}_{21}=\tau(\mbox{$\cal R$})$ and $\tau$ is the abstract permutator. Since this equation holds for an arbitrary representation $\rho$, we can drop the latter symbol and obtain the \paragraph{Universal expression for the twisted permutation operator of $U_q(su(2))$:} \begin{equation} P_{12}^{\scriptscriptstyle {\cal F}_{12}}=f({\bf 1}\otimes C_q)f(C_q\otimes {\bf 1})\left[f( \Delta(C_q))\right]^{-1}\mbox{$\cal R$}_{21}\circ \tau \end{equation} We omit here the well-known expression for the universal $\mbox{$\cal R$}$ \cite{dr}. \subsection{$n\ge 3$ particles} When $n\ge 3$, for any given space $V$ the decomposition of $\bigotimes^n V$ into irreducible representations of the permutation group contains components with partial/mixed symmetry, beside the completely symmetric and the completely antisymmetric ones. \footnote{The Young tableaus provide the rules for finding the complete decomposition for any $n$.} If $n=3$, for instance, some components can be diagonalized {\it either} w.r.t. to $P_{12}$ {\it or} w.r.t. $P_{23}$ (but not w.r.t. both of them simultaneously). If $n=4$, all components can be diagonalized simultaneously w.r.t. $P_{12}$ and $P_{34}$, and some will have mixed symmetry (e.g. will be symmetric in the first pair and antisymmetric in the second, or vice versa). We recall that the explicit knowledge of components with mixed/partial symmetry is required to build $({\cal H}^{\otimes n})_{\pm}$ if the Hilbert space ${\cal H}$ of one particle is the tensor product of different spaces, ${\cal H}=V\otimes V'$, as in example~2 in subsection~\ref{transta}. It is easy to realize that similar statements hold in the case of the twisted symmetry. Let us consider again the case $V_j$, and let $n=3$ for the sake of simplicity. We show how to construct two different orthonormal bases of $V_j\otimes V_j\otimes V_j$ with (partial) symmetry, and a continuous family of $F_{123}$ on $V_j\otimes V_j\otimes V_j$. There is evidently only one irreducible representation with highest weight $J=3j$, the highest weight vector being $|j,j\rangle|j,j\rangle|j,j\rangle$. But there are two independent irreducible representations with highest weight $J=3j-1$, e.g. those having highest weight vectors $\frac1{\sqrt{2}}\left(|j,j-1\rangle|j,j\rangle\pm|j,j\rangle |j,j-1\rangle\right)|j,j\rangle$. The latter are symmetric and antisymmetric respectively w.r.t. $P_{12}$, but are mixed into each other by the action of $P_{23}$; alternatively, one can combine these two representations into two new ones, having highest weight vectors $\frac1{\sqrt{2}}|j,j\rangle\left(|j,j-1\rangle|j,j\rangle\pm|j,j\rangle |j,j-1\rangle\right)$, which are symmetric and antisymmetric respectively w.r.t. $P_{23}$, but are mixed into each other by the action of $P_{12}$. One can easily verify that the first two representations are eigenspaces of $\rho^{(2)}_c(C_c)\otimes id$ with eigenvalues $(2j\pm \frac 12)^2$, the latter two are eigenspaces of $id \otimes \rho^{(2)}_c(C_c)$ with the same eigenvalues. The operators $\rho^{(3)}_c(C_c),\rho^{(3)}_c(h)$ and either $\rho^{(2)}_c(C_c)\otimes id$ or $id \otimes \rho^{(2)}_c(C_c)$ make up a complete set of commuting observables over $V_j\otimes V_j\otimes V_j$. Let \begin{equation} \{\Vert J,M,r\rangle_{12}\}_{J,M,r}, \qquad [\mbox{respectively:}\quad \{\Vert J,M,s\rangle_{23}\}_{J,M,s}], \end{equation} with \begin{equation} j\le J\le 3j, \hspace*{9mm} -J\le M\le J,\hspace*{9mm} \mbox{max}\{0,j\!-\!J\}\le r,s\le \mbox{min}\{2j,j\!+\!J\} \label{prima} \end{equation} denote an orthonormal basis of eigenvectors of $\rho^{(3)}_c(C_c)$, $\rho^{(3)}(h)$ and $\rho^{(2)}_c(C_c)\otimes id$ [respectively:\hspace{3mm} $id \otimes \rho^{(2)}_c(C_c)$] with eigenvalues $J(J+1),M$ and $r(r+1)$ [respectively:\hspace{3mm} $s(s+1)$]. In particular, \begin{eqnarray} \Vert 3j-1,3j-1,2j-\frac 12 \pm\frac 12 \rangle_{12} = \frac1{\sqrt{2}}\left(|j,j-1\rangle|j,j\rangle\pm|j,j\rangle |j,j-1\rangle\right) |j,j\rangle\nonumber\\ \Vert 3j-1,3j-1,2j-\frac 12 \pm\frac 12 \rangle_{23} = \frac1{\sqrt{2}}|j,j\rangle\left(|j,j-1\rangle|j,j\rangle\pm|j,j\rangle |j,j-1\rangle\right) \end{eqnarray} It is easy to verify that in general the subspace of $V_j\otimes V_j\otimes V_j$ which is anti-symmetric/symmetric w.r.t.\ $P_{12}$ is spanned by the vectors $\Vert J,M,r\rangle_{12}$ with $r-\mbox{min}\{2j,j\!+\!J\}$ odd/even, and similarly for $P_{23}$. For fixed $J,M$, there exists a unitary matrix $U(J)$ such that \begin{equation} \Vert J,M,s\rangle_{23}=U(J)_{sr}\Vert J,M,r\rangle_{12} \label{ultima} \end{equation} Formulae formally identical to eqs. (\ref{prima}), (\ref{ultima}) hold when $q\neq 1$; we will introduce an additional index $q$ in all objects to denote this dependence. The elements of $\rho^{(3)}(U_q(su(2)))$ in these two bases read \begin{equation} \rho^{(3)}(X)=\cases{\sum\limits_J\sum\limits_r\sum\limits_{M,M'} X_{M,M'}(J) \Vert J,M,r,q\rangle_{12}~ _{12}\langle J,M,r,q\Vert\cr \sum\limits_J\sum\limits_s\sum\limits_{M,M'} X_{M,M'}(J) \Vert J,M,s,q\rangle_{23} ~_{23}\langle J,M,s,q\Vert, \cr} \end{equation} and the matrix elements $X_{M,M'}(J)$ do not depend on $r,s$. Now it is easy to check that we can find many-parameter continuous families of matrices $F_{123}$ satisfying eq. (\ref{condition}), in the form \begin{equation} F_{123}=\cases{\sum\limits_J\sum\limits_M\sum\limits_r A_{r,r'}(J) \Vert J,M,r,q\rangle_{12}~_{12}\langle J,M,r',1\rangle\Vert \cr \sum\limits_J\sum\limits_M\sum\limits_s B_{s,s'}(J) \Vert J,M,s,q\rangle_{23} ~_{23}\langle J,M,s',1\Vert, \cr} \label{family} \end{equation} where $A(J)$'s are arbitrary unitary matrices and $B(J)=[U(J,q)]^*A(J) [U(J,q=1)]^T$. The key point is that the matrix elements $A_{r,r'}$ do not depend on $M$, whereas the matrix elements $X_{M,M'}$ do not depend on $r$. It is easy to realize that the family (\ref{family}) interpolates between the two $F$ matrix given in subsection \ref{transta}, $F_{123}'$ (if we set $A_{r,r'}=\delta_{r,r'}$) and $F_{123}''$ (if we set $B_{r,r'}=\delta_{r,r'}$). Considerations analogous to those of subsection \ref{example1} can be done for $n\ge 3$ when $V$ is a reducible representation of $U_q(su(2))$. \section*{Acknowledgments} We would like to thank R.\ Engeldinger for sharing his interpretation of Drinfel'd twists which helped us in a change of viewpoint on the problem and we are grateful to S.\ L.\ Woronowicz and W.\ Pusz for pointing out Ref. \cite{wopu}. One of us (G.F.) thanks A.~v.~Humboldt foundation for financial support. Last but not least we thank J.~Wess for discussion and warm hospitality at his institute.
\section{Introduction} \setcounter{equation}{0} \setcounter{page}{2} The Casimir effect \cite{casimir48} is a beautiful and simple manifestation of the influence that boundaries or non-trivial spacetime topologies have on quantum field theories (see for example \cite{plunienmullergreiner86,ambjornwolfram83,eorbz,moste}). The modern approach for the calculation of the Casimir energy is the zeta function regularization scheme put forward in \cite{blauvisserwipf88}. The idea is the following. In order to have a well-defined notion of energy, let us work in a $d$-dimensional ultrastatic spacetime \cite{dewitt75}. Thus the metric in some coordinate system is of the form $g=-(dx^0)^2 +g_S$ with the spatial part $g_S$ of the metric (see Eq.~(\ref{eq:2.1})). The differential operator $D$ describing the field equation may be decomposed as $D=-\partial _0^2 +D_S$. Introducing $E_n^2=\lambda_n$ as the eigenfrequencies of $D_S$, the zero-point energy is formally given by \begin{equation} E_{Casimir} =\frac 1 2 \sum_n E_n. \label{eq:1.1} \end{equation} It may be regularized by defining \begin{equation} E_{reg}(\epsilon ) =\frac 1 2 \mu^{2\epsilon} \zeta _S (-1/2 +\epsilon ) , \label{eq:1.2} \end{equation} with $\zeta _S (s)$ being the zeta function associated with the ($d-1$)-dimensional operator $D_S$. The scale $\mu$ with dimension (length)$^{-1}$ has to be introduced in order to keep the zeta function dimensionless for all $s$. General zeta function theory \cite{voros87} tells us, that $E_{reg} (\epsilon ) $ is a meromorphic function with a pole at $\epsilon =0$, its residue being $-(1/2) C_{d/2} (D_S)/(4\pi )^{d/2}$. The coefficient $C_{d/2} (D_S)$ is the Seeley-De Witt coefficient appearing in the asymptotic expansion for small $t$ of the heat-kernel associated with $D_S$, \begin{eqnarray} K(t) &=& \sum_n e^{-\lambda_n t}\nonumber\\ &\sim & \left(\frac 1 {4\pi t}\right)^{(d-1)/2} \sum_{l=0,1/2,1,...} ^{\infty}C_l(D_S) t^l .\label{eq:1.3} \end{eqnarray} The pole appearing in Eq. (\ref{eq:1.2}) has to be absorbed into the bare action which thus must contain a term proportional to $C_{d/2} (D_S)$. It is clear then, that the Casimir energy has an ambiguity proportional to $C_{d/2} (D_S)$ (we will come back to this point later). Adopting the minimal subtraction scheme, one defines \begin{eqnarray} E_{Casimir} &=& \frac 1 2 \lim_{\epsilon \to 0} \frac 1 2 \mu^{2\epsilon}\left[\zeta_S (-1/2 +\epsilon ) +\zeta_S (-1/2-\epsilon )\right]\nonumber\\ &=& \frac 1 2 \left[ PP\,\, \zeta_S (-1/2)-\frac{C_{d/2}(D_S) }{(4\pi)^{d/2}}\ln\mu^2\right] , \label{eq:1.4} \end{eqnarray} where the symbol $PP$ stands for taking the principal part. Based on this definition, during the last years the Casimir energy has been calculated for a variety of examples \cite{eorbz,dolannash92,carusonetosvaitersvaiter91}. The definition (\ref{eq:1.4}) of the Casimir energy is completely equivalent to the one steming naturally from the definition of functional determinant by the zeta function prescription (for very recent considerations on this issue see \cite{zds}). This may be done by considering the theory at finite temperature and by defining the Casimir energy as its energy in the limit $T\to 0$, this idea going back to Gibbons \cite{gibbons77} who considered the single quantum mechanical oscillator. In the more general context of quantum field theory under some external conditions like boundaries or gravitational fields this definition has been employed for example in \cite{dowkerkennedy78,ambjornwolfram83,dowker84,cognolavanzozerbini92,trento}). The ambiguity in the coefficient $C_{d/2}$ may be understood to be a result of the necessary renormalization of the free energy of the system. Having summarized the main arguments in favor of the definition of the Casimir energy as given by Eq.~(\ref{eq:1.4}), in section 3 we will present the calculation of the Casimir energy for a massive scalar field in a general $(1+2)$-dimensional toroidal spacetime with flat spatial geometry. The general flat geometry will be parametrized by the corresponding two Teichm\"uller parameters and the complete dependence of the Casimir energy on these parameters and on the mass of the field will be obtained under the form of an analytic function, by using the extended Chowla-Selberg zeta function formula derived by one of us in Refs.~\cite{eecs1,eecs2}. For the massless case we obtain complete agreement with previous results by Dowker \cite{dowker89a}. \section{Zeta function definition of the Casimir energy} \setcounter{equation}{0} Let us first briefly summarize the motivation for the definition (\ref{eq:1.4}) of the Casimir energy. For definiteness, let us consider the quantum field theory of a free scalar field in curved spacetime, the Dirac field may be treated analogously. As mentioned before, in order to have a well defined notion of energy we shall restrict our considerations to a $d$-dimensional ultrastatic spacetime ${\cal M}$, possibly with a boundary, and with the metric \begin{eqnarray} ds^2=-dx_0^2+g_{ab}(\vec x )dx^adx^b,\label{eq:2.1} \end{eqnarray} where $\vec x =(x_1,...,x_{d-1})$. The action of the field theory we consider is \cite{dewitt75}, \cite{critchleydowkerkennedy80}, \begin{eqnarray} S=-\frac 1 2 \int\limits_{{\cal M}} d^d x |g|^{\frac 1 2}\phi^{\dagger}(x) \left(\Box -\xi R-m^2\right)\phi (x),\label{22} \end{eqnarray} with $\Box$ being the Laplace-Beltrami operator of the ultrastatic spacetime. Variation of equation (\ref{22}) subject to the constraints \begin{eqnarray} \delta \phi (x')&=&0,\nonumber\\ n^{\mu '}\nabla _{\mu '}\delta \phi (x')&=&0,\nonumber \end{eqnarray} where the prime refers to quantities defined on the boundary $\partial {\cal M}$, yields the equation of motion \begin{eqnarray} \left(\Box -\xi R-m^2\right) \phi (x)=0,\label{25} \end{eqnarray} which is the generalized Klein-Gordon equation. The following discussion will be quite general, so the boundary condition need not to be specified at this point. A unique boundary value problem is posed, for example, by assuming Dirichlet- or Robin-boundary conditions on the field. In an ultrastatic spacetime the quantum field theory at finite temperature may be developed in complete analogy with the Minkowski-space theory and that is why we will skip the details of the calculation. In the Euclidean formulation of the finite temperature theory, the partition sum ${\cal Z}$ may be written under the form of a functional integral of the kind \begin{equation} {\cal Z} [\beta ] =\int [d\varphi]\exp\left\{-\frac 1 2 (\varphi ,D\varphi )\right\}, \end{equation} where we have used the scalar product \begin{equation} (f,h) =\int\limits_0^{\tau} d\tau \int\limits_{\Sigma}d\Sigma\,\, |g|^{1/2} f^* h\nonumber \end{equation} for the two vectors $f$ and $h$, and being $\Sigma$ the spatial section of the manifold ${\cal M}$. Here the integration extends over all fields periodic in the imaginary time $\tau$ with periodicity $\beta =1/T$ and fulfilling the boundary conditions at the spatial boundary. The operator $D$ is given by \begin{equation} D =-\frac{\partial^2}{\partial \tau^2} -\Delta +\xi R +m^2. \end{equation} Then, one formally obtains \begin{equation} {\cal Z} [\beta , \mu ] =(\det \lambda ^{-2} D)^{-\frac 1 2}, \end{equation} the scale $\lambda$ \cite{hawking77} being necessary in order to keep everything dimensionless. The functional determinant of the operator $\lambda^{-2} D$ needs, of course, regularization. We will use the zeta-function regularization scheme introduced by Dowker, Critchley \cite{critchleydowker76} and Hawking \cite{hawking77}. In this scheme, the free energy is defined as \begin{equation} F[\beta ]=-\frac 1 {2\beta} \left[\zeta_d (0,\beta )\ln \lambda^2 +\zeta_d ' (0,\beta )\right].\label{free} \end{equation} The function $\zeta_d(s,\beta)$ is the zeta-function associated with the operator $D$. Using the ansatz \begin{eqnarray} u_{l,k}=\frac 1 {\beta} \exp\left(\frac{2\pi il}{\beta}\right)g_k(\vec x)\nonumber \end{eqnarray} the eigenvalues are seen to have the form \begin{eqnarray} \nu_{l,k}^{\pm}=\left(\frac{2\pi l}{\beta}\right)^2+E_k^2,\,\, \ \ l\in{\mbox{${\rm Z\!\!Z }$}},\nonumber \end{eqnarray} with the energy eigenvalues being defined through \begin{equation} (-\Delta +\xi R +m^2) \psi_k (\vec x )=E_k^2 \psi_k (\vec x). \end{equation} Making use of a Mellin-transformation and a theta-function identity \cite{hille62}, the free energy may be written in the form \begin{eqnarray} F[\beta]&=&\frac 1 2 PP\zeta_S\left(-\frac 1 2\right)+\frac 1 {2(4\pi)^{\frac d 2}}C_{\frac d 2}[\ln \lambda^2-1+2\ln 2]\label{endre}\\ &+&\frac 1 {\beta} \sum_j\ln\left(1-e^{-\beta E_j}\right)\nonumber. \end{eqnarray} The energy of the system is then given by \begin{eqnarray} E=\frac{\partial}{\partial\beta}\beta F[\beta ]. \label{energie} \end{eqnarray} Finally, defining the Casimir-energy as the limit of (\ref{energie}) for $T\to 0$, we obtain \begin{eqnarray} E_{Cas}[\partial{\cal M}]=\lim_{T\to 0}E=\frac 1 2 PP\zeta_S\left(-\frac 1 2\right)+\frac 1 {2(4\pi)^{\frac d 2}}C_{\frac d 2}\ln \tilde{\lambda}^2,\label{casimir} \end{eqnarray} with $\tilde{\lambda}=\frac{2\lambda}{\sqrt{e}}$. We thus arrive to the definition of Refs.~\cite{dowker84,blauvisserwipf88}. In the last reference a detailed discussion of the meaning of this definition and of the problem of renormalization has been carried out. From the above derivation of the definition of the Casimir energy it is completely clear that the ambiguity of the Casimir energy is simply a result of the (in general) necessary renormalization of the free energy. In the case when the coefficient $C_{d/2}$ vanishes, the definition (\ref{casimir}) gives a well defined, finite value. \section{Casimir energy in a (1+2)-dimensional toroidal spacetime} \setcounter{equation}{0} Let us consider, as an example, a (1+2)-dimensional spacetime with the topology $\mbox{${\rm I\!\!R }$} \times T^2$ \cite{seriu}. We will concentrate on the case when the geometry of the space $\Sigma\simeq T^2$ is locally flat. One can construct such geometry, as is usual, by taking $\Sigma =[0,1]\times [0,1]/\sim $, where the equivalence relation is defined by $(\xi_1 ,0) \sim (\xi_1 ,1)$ and $ (0,\xi _2 )\sim (1,\xi_2)$. A flat $2$-geometry is endowed on $\Sigma$ by giving it a metric \begin{equation} ds^2 =h_{ab} d\xi^a d\xi^b, \end{equation} where \begin{equation} h_{ab}=\frac 1 {\tau_2}\left( \begin{array}{cc} 1 & \tau_1\\ \tau_1 & |\tau|^2 \end{array}\right) . \end{equation} The $(\tau_1, \tau_2)$ are the Teichm\"uller parameters, independent of the spatial coordinates $(\xi_1,\xi_2)$, and $\tau =\tau_1+i\tau_2$, $\tau_2>0$ \cite{hatfield92,luesttheisen89}. The Laplace-Beltrami operator of this metric is given by \begin{equation} \Delta =-\frac 1 {\tau_2} (|\tau|^2 \partial_1^2 -2\tau_1\partial_1\partial_2 +\partial_2^2 ), \end{equation} being its eigenvalues \begin{equation} \lambda_{n_1,n_2} =\frac{4\pi^2}{\tau_2} (|\tau|^2n_1^2 -2\tau_1n_1n_2 +n_2^2 ). \end{equation} In the massive case, $m \neq 0$ the spectrum runs over $n_1,n_2 \in \mbox{${\rm Z\!\!Z }$}$. In the massless case the zero-mode of $\Delta$, $n_1=n_2=0$, has to be excluded. An exact analysis of the Casimir energy for this spacetime is possible since it reduces to a case of the Chowla-Selberg zeta function (when $m=0$) or to one of the extended formula that has been obtained recently (case $m\neq 0$). In fact, in the case when $m \neq 0$, the relevant formula is a particular application of the following. Let us consider the double series \begin{equation} E(s;a,b,c;q) \equiv {\sum_{m,n \in \mbox{\bf Z}}}' (am^2+bmn+cn^2+q)^{-s}, \label{1} \end{equation} with $q\neq 0$ (in general), the parenthesis in (\ref{1}) is the inhomogeneous quadratic form \begin{equation} Q(x,y)+q, \ \ \ \ Q(x,y) \equiv ax^2+bxy+cy^2, \end{equation} restricted to the integers. In the general theory that deals with the homogeneous case, one assumes that $a,c >0$ and that the discriminant \begin{equation} \Delta =4ac-b^2 >0 \end{equation} (see \cite{cs}). Here we will impose the additional condition that $q$ be such that $Q(m,n)+q \neq 0$, for all $ m,n \in $ {\bf Z}. In the usual applications of the theory, those conditions are indeed satisfied. For the analytical continuation of (\ref{1}), the following expression has been obtained in Refs. \cite{eecs1,eecs2} \begin{equation} E(s;a,b,c;q) = 2\zeta_{EH} (s,q/a)\, a^{-s} + \frac{2^{2s} \sqrt{\pi}\, a^{s-1}}{\Gamma (s) \Delta^{s-1/2}} \, \Gamma (s - 1/2) \zeta_{EH} (s-1/2,4aq/\Delta) \ \ \ \ \ \ \ \ \label{gcs} \end{equation} \[ + \frac{2^{s+5/2} \pi^s }{\Gamma (s) \sqrt{a}} \sum_{n=1}^\infty n^{s-1/2} \cos (n \pi b/a) \sum_{d|n} d^{1-2s} \left( \Delta + \frac{4aq}{d^2} \right)^{-s/2+1/4} K_{s - 1/2}\left( \frac{\pi n}{a} \sqrt{ \Delta + \frac{4aq}{d^2}} \right), \] $\sum_{d|n}d^s$ denoting the sum over the divisors of $n$ and where the function $\zeta_{EH} (s,p) $ (the one dimensional Epstein-Hurwitz or inhomogeneous Epstein series) is given by \begin{eqnarray} \zeta_{EH}(s;p) &=& \sum_{n=1}^\infty \left( n^2 + p \right)^{-s} = \frac{1}{2} {\sum_{n \in \mbox{\bf Z}}}' \, \left( n^2 + p \right)^{-s} \label{zeh1} \\ & =& -\frac{p^{-s}}{2} + \frac{\sqrt{\pi} \, \Gamma (s- 1/2)}{2\, \Gamma (s)} p^{-s+1/2} + \frac{2\pi^s p^{-s/2 +1/4}}{\Gamma (s)} \sum_{n=1}^\infty n^{s -1/2} K_{s -1/2} (2\pi n\sqrt{p}). \nonumber \end{eqnarray} Eq. (\ref{gcs}) provides the analytical continuation of the inhomogenous Epstein series, in the variable $s$, as a meromorphic function in the complex plane. Its pole structure is explicitly given in terms of the well-known pole structure of $\zeta_{EH} (s,p)$. Eq. (\ref{gcs}) has been found by one of us and called the {\it extended Chowla-Selberg} formula, since it contains the Chowla-Selberg formula as the particular case $q=0$, i.e., \begin{eqnarray} && E(s;a,b,c;0) = 2\zeta (2s)\, a^{-s} + \frac{2^{2s} \sqrt{\pi}\, a^{s-1}}{\Gamma (s) \Delta^{s-1/2}} \,\Gamma (s -1/2) \zeta (2s-1) + \frac{2^{s+5/2} \pi^s }{\Gamma (s) \Delta^{s/2-1/4}\sqrt{a}} \nonumber \\ && \hspace{1cm} \times \sum_{n=1}^\infty n^{s-1/2} \, \sigma_{1-2s} (n) \, \cos (n \pi b/a) \, K_{s - 1/2}\left( \frac{\pi n \sqrt{\Delta}}{a} \right). \label{cs1} \end{eqnarray} where \begin{equation} \sigma_s(n) \equiv \sum_{d|n} d^s. \end{equation} Formula (\ref{gcs}) has been obtained for the first time in \cite{eecs1}, and a misprint has been corrected in \cite{eecs2}. The good convergence properties of expression (\ref{cs1}), that were so much prised by Chowla and Selberg, are shared by its non-trivial extension (\ref{gcs}). This renders the use of the formula quite simple. In fact, the two first terms are just nice ---under the form (\ref{zeh1})--- while the last one (impressive in appearence) is even more quickly convergent than in the case of Eq. (\ref{cs1}), and thus absolutely harmless in practice. Only a few first terms of the three series of Bessel functions in (\ref{gcs}), (\ref{zeh1}) need to be calculated, even if one demands good accuracy. We should also notice that the only pole of (\ref{cs1}) at $s=1$ appears through $\zeta (2s-1)$ in the second term, while for $s=1/2$, the apparent singularities of the first and second terms cancel each other and no pole is formed. Analogously, the pole at $s=1/2$ in (\ref{gcs}) comes only from the first term. Eq. (\ref{gcs}) also has these good properties, {\it for any non-negative value of} $q$. In fact, for large $q$ the convergence properties of the series of Bessel functions are clearly enhanced, while for $q$ small we get back to the case of Chowla and Selberg. Notice, however, that this is not obtained through the high-$q$ expansion (e.g., just putting $q=0$ in (\ref{gcs})), but using a low-$q$, binomial expansion of the kind \begin{equation} \sum_{n=0}^{\infty} \left[ a(n+c)^2+q\right]^{-s}= a^{-s} \sum_{m=0}^{\infty} \frac{(-1)^m \Gamma(m+s)}{\Gamma (s) \, m!} \left( \frac{q}{a} \right)^m \zeta_H (2s+2m,c), \end{equation} which is convergent for $q/a \leq 1$. For $q \rightarrow 0$ it reduces to $a^{-s} \zeta_H (2s,c)$. Actually, formula (\ref{gcs}) is still valid in a domain of negative $q$'s, namely for $q > -\min (a,c,a-b+c)$. Turning now to the particular application of the formula in our specific situation, we see that the zeta function corresponding to the Laplace-Beltrami operator in the massive (resp. massless) case is simply given by: \begin{equation} \zeta_{\Delta + m^2} (s) = m^{-2s} + \left( \frac{4\pi^2}{\tau_2} \right)^{-s} E\left(s; |\tau|^2, -2\tau_1,1; \frac{\tau_2m^2}{4\pi^2} \right) \end{equation} and \begin{equation} \zeta_{\Delta} (s) = \left( \frac{4\pi^2}{\tau_2} \right)^{-s} E\left(s; |\tau|^2, -2\tau_1,1; 0 \right), \end{equation} respectively. The values at $s=-1/2$ are finite and define the corresponding Casimir energy. After performing the necessary calculations, and according to the prescription that has been derived in the first part of this paper, we obtain as result the quite simple expressions \begin{eqnarray} && \zeta_{\Delta + m^2} (-1/2) = -\frac{m^3}{6\pi} - \frac{2m}{\pi} \sum_{n=1}^\infty n^{-1} K_1 (nmx_2) - \sqrt{2} \left( \frac{m x_2}{\pi}\right)^{3/2} \sum_{n=1}^\infty n^{-3/2} K_{3/2} (nm/x_2) \nonumber \\ && -8 x_2\sum_{n=1}^\infty n^{-1} \cos (2\pi n x_1^2) \sum_{d|n} d^{2} \sqrt{ 1 + \frac{m^2}{(2\pi x_2 d)^2}} \ K_1 \left( 2\pi n x_2^2 \sqrt{ 1 + \frac{m^2}{(2\pi x_2 d)^2}} \right), \label{d12m} \end{eqnarray} and \begin{equation} \zeta_{\Delta } (-1/2) = -\frac{\pi}{3x_2} + 4\pi \zeta'(-2) x_2^3 -8 x_2\sum_{n=1}^\infty n^{-1} \cos (2\pi n x_1^2) \sigma_2(n) \, K_1 \left( 2\pi n x_2^2 \right), \label{d12} \end{equation} in terms of the variables \begin{equation} x_1 = \sqrt{\frac{\tau_1}{\tau_1^2 + \tau_2^2}}, \qquad x_2 = \sqrt{\frac{\tau_2}{\tau_1^2 + \tau_2^2}}. \label{vch} \end{equation} The extrema of the corresponding Casimir energy for the case $m^2=100$, in terms of the original Teichm\"uller coefficients $\tau_1$ and $\tau_2$, are to be read from Figs. 1-3. In the three-dimensional plot over the plane $\tau_1, \tau_2$ (Fig. 1), the maximal Casimir energy is seen to be localized on the section $\tau_2=1$. In order to show this fact more clearly, in Fig. 2 we have represented the section $\tau_1=0$, but the situation is common to any section $\tau_1=$ const. On the section $\tau_2=1$ a periodic structure appears asociated with the value of $\tau_1$ along this section (Fig. 3). This behavior is easy to recognize from the form of the function $\zeta_\Delta (-1/2)$, (\ref{d12}), and is common to {\it any} section $\tau_2=$ const. All the figures depicted here have been obtained taking the first 20 terms from the sum over $n$ in (\ref{d12}). \vspace{5mm} \noindent{\large \bf Acknowledgments} We thank all the members of the Department of Theoretical Physics of the University of Trento for warm hospitality, specially Sergio Zerbini, Luciano Vanzo, Guido Cognola, Ruggero Ferrari and Marco Toller. Thanks are given to the referee of a previos version of this paper for very precise comments that led to its improvement. This work was finished while EE was visiting the Institute of Theoretical Physics, Chalmers University of Technology (Sweden), and has been supported by DGICYT and Ministry of Foreign Affairs (Spain), by CIRIT (Generalitat de Catalunya) and by INFN (Italy). \newpage
\section{\bf Introduction} \hspace{0.5cm} A heavy fermion superconducting $U\!Pt_{3}$ compound is an example of the unconventional superconductivity, in which both the gauge and the point group symmetries are broken in the ordered phase. At the temperature $T_{N}\simeq 5K$ it undergoes the antiferromagnetic transition with the magnetic moments confined to the $D_{6h}$ basal plane, however the long range antiferromagnetic correlations have not yet been seen \cite{1,2}. Far below the N\'eel temperature, at $T_{c_{+}}\simeq 0.51K$ (p=0 bar) $U\!Pt_{3}$ becomes superconducting \cite{3,4}. There is another superconducting transition at $T_{c_{-}}\simeq 0.46K$ (p=0 bar) \cite{3,4}. This feature and a rich phase diagram in the magnetic field and temperature plane \cite{5} are accepted as the evidence of a multicomponent superconducting order parameter. There are also the pressure experiments which indicate strongly the coupling between superconductivity and magnetism in $U\!Pt_{3}$ \cite{2,7,8}. Namely the specific heat measurements under pressure show that the two critical temperatures $T_{c_{+}}$ and $T_{c_{-}}$ converge into one critical temperature $T_{c}$ above $p_{c}\simeq 4$ kbar pressure value Fig.1 \cite{2,7,8}, which is the pressure that destroyes antiferromagnetism in the system. This experiment supports the theory of a two component order parameter $\bar{\psi}=(\psi_{x},\psi_{y})$ in a basal plane of the crystal, belonging to a two-dimensional irreducible representation of the hexagonal point group $D_{6h}$. In this approach a complex vector $\bar{\psi}$ couples to the magnetic moment $\bar{M}$ and the split transition is due to that interaction. The role of magnetism as a symmetry breaking field coupling to superconductivity is revealed in the neutron scattering measurements \cite{1,10}. In these experiments Aeppli et al. established that below the temperature of the order of a superconducting transition temperature the neutron scattering intensity of the $(1,\frac{1}{2},0)$ reflection suddenly saturates and is almost constant unless the superconductivity occurs. There is a remarkable change in the temperature dependence for a superconducting system. At a temperature of the order of $T_{c}$ the slope of the neutron scattering intensity changes sign and the intensity becomes an increasing function of temperature see Fig.2 \cite{1,10}. This is another strong evidence of the coupling between magnetism and superconductivity in $U\!Pt_{3}$.\\ \hspace*{0.5cm} Recently Joynt \cite{11} discussed within a two component order parameter approach the phase diagram of $U\!Pt_{3}$ in three-dimensional magnetic field-pressure-temperature space. It agrees qualitatively with measurements \cite{2,5,7,8,12}. However the temperature dependence of the magnetic moment observed by Aeppli et al. \cite{1,10} was not taken into account. The contradiction here arises as follows. The magnetic Bragg peak observed in neutron scattering \cite{1,10} reproduced in Fig.2 shows that the superconductivity is acting to suppress the magnetism. By thermodynamic reasoning we know that if the onset of superconductivity reduces the magnetism then the onset of magnetism must reduce the tendency to superconductivity. The magnetism may be removed by pressure \cite{2,7,8}. We observe that as the pressure is reduced below $p_{c}$ where the magnetism reappears the slope of transition temperature is increased see Fig.1. In other words the critical temperatures $T_{c_{+}}$ and $T_{c_{-}}$ are not suppressed equally by the pressure, what can be expressed quantitatively by an inequality which follows : \begin{eqnarray} \label{e0} \frac{T_{c_{+}}\left(p=0\right)-T_{c}\left(p=0\right)} {T_{c}\left(p=0\right)-T_{c_{-}}\left(p=0\right)} & > & 1\;, \end{eqnarray} see Fig.1.\\ We show that this competition between superconductivity and antiferromagnetism cannot be understood within the simple model of magnetism considered so far.\\ \hspace*{0.5cm} The plan of the paper is as follows. In section 2 we study mentioned already pressure and magnetic experiments in a frame of \mbox{two-dimensional} superconducting order parameter scenario. To avoid the inconsistencies following from this approach we introduce a two magnetic moment model in section 3. Within this scenario we analyse the experimental data and obtain several constraints on the \mbox{Ginzburg-Landau} free energy coefficients in sections 3 and 4. Finally we summarize the results in section 5. \section{\bf Two component superconductivity coupled to magnetism} \hspace{0.5cm} In this section we review the experimental evidence which supports this model and then construct the free energy. The free energy is used to obtain the coupled order parameters of magnetism and superconductivity. This analysis follows \cite{9,16,17}. We reproduce it here because it is important to consider both the temperature and pressure experiments using a unified notation. In this approach we start with a free energy density : \begin{equation} \label{e1} F=F_{M}+F_{S}+F_{SM}, \end{equation} where \begin{equation} \label{e2} F_{M}=\left\{ \begin{array}{ccl} \alpha_{M}(T^{\ast}-T_{N})M^{2}+\frac{1}{2}\beta_{M}M^{4}, & for & T \leq T^{\ast}\\ \alpha_{M}(T-T_{N})M^{2}+\frac{1}{2}\beta_{M}M^{4}, & for & T > T^{\ast} \end{array} \right. \end{equation} \begin{equation} \label{e3} F_{S}=\alpha_{S}(T-T_{c})|\bar{\psi}|^{2}+\frac{1}{2}\beta_{1}|\bar{\psi}|^{4}+ \frac{1}{2}\beta_{2}|\bar{\psi}^{2}|^{2}, \end{equation} \begin{equation} \label{e4} F_{SM}=\gamma|\bar{M}\bar{\psi}|^{2}+\alpha M^{2}|\bar{\psi}|^{2}. \end{equation} All the Ginzburg-Landau coefficients are very weakly temperature and pressure dependent, what can be shown in a weak-coupling microscopic theory \cite{14}, hence we choose them to be constant. The magnetic free energy given by equation $\mbox{(\ref{e2})}$ has been chosen to include the phenomenological saturation of $M$ below $T^{\ast}$ \cite{1,10}. The coefficients in $F_{M}\; \mbox{(\ref{e2})}$ and $F_{S}\;\mbox{(\ref{e3})}$ are positive whereas the $\gamma$ coefficient in $F_{SM}\;\mbox{(\ref{e4})}$ may be chosen to be negative and $\bar{M}$ is then parallel to $\hat{x}$. The superconducting order parameter $\bar{\psi}=(\psi_{x},\psi_{y})$ is complex and its composits $\psi_{x}$ and $\psi_{y}$ are written as \mbox{$\psi_{x}=|\psi_{x}|{\mit e}^{{\mit i}\varphi_{x}}$} and \mbox{$\psi_{y}=|\psi_{y}|{\mit e}^{{\mit i}\varphi_{y}}$.}\\ \hspace*{0.5cm}Minimisation of the free energy leads to the following equations for the order parameters : \begin{equation} \label{e01} 0=\alpha_{M}(T_{M}-T_{N})+\beta_{M}M^{2}+\gamma|\psi_{x}|^{2}+ \alpha|\bar{\psi}|^{2}\;\;\;{\rm or}\;\;\;M=0\;, \end{equation} where \begin{displaymath} T_{M}=\left\{ \begin{array}{ccl} T\;, & for & T>T^{\ast}\\ T^{\ast}\;, & for & T\leq T^{\ast}\;, \end{array} \right. \end{displaymath} \begin{equation} \label{e02} 0=\alpha_{S}(T-T_{c})+\beta_{1}|\bar{\psi}|^{2}+\beta_{2}(|\psi_{x}|^{2}+ |\psi_{y}|^{2}\cos 2(\varphi_{x}-\varphi_{y}))+\gamma M^{2} + \alpha M^{2} \end{equation} \hspace*{0.3cm} or $\;\;\;|\psi_{x}|=0$, \begin{equation} \label{e03} \left\{ \begin{array}{l} 0=\alpha_{S}(T-T_{c})+\beta_{1}|\bar{\psi}|^{2}+\beta_{2}(|\psi_{y}|^{2}+ |\psi_{x}|^{2}\cos 2(\varphi_{x}-\varphi_{y}))+\alpha M^{2}\\ {\rm and}\;\;\varphi_{x}-\varphi_{y}=\frac{\pi}{2} \end{array} \right. \end{equation} \hspace*{0.5cm} or $\;\;\;|\psi_{y}|=0$.\\ \vspace{0.3cm}\\ \hspace*{0.5cm} From these expressions we find the following conditions for $M$, $\psi_{x}$ and $\psi_{y}$ : \begin{equation} \label{e04} M=|\psi_{x}|=|\psi_{y}|=0\;\;\;\;for\;\;\;\;T>T_{N}\;, \end{equation} \begin{equation} \label{e05} M^{2}=\frac{\alpha_{M}}{\beta_{M}}(T_{N}-T),\;\;\;\;|\psi_{x}|=|\psi_{y}|=0 \;\;\;\;for\;\;\;\;T^{\ast}<T \leq T_{N}\;, \end{equation} \begin{equation} \label{e06} M^{2}=\frac{\alpha_{M}}{\beta_{M}}(T_{N}-T^{\ast}),\;\;\;\;|\psi_{x}|= |\psi_{y}|=0\;\;\;\;for\;\;\;\;T_{c_{+}}<T\leq T^{\ast}\;, \end{equation} \begin{equation} \label{e07} \left\{ \begin{array}{l} M^{2}=\frac{1}{\beta_{M}}\left[\alpha_{M}(T_{N}-T^{\ast})-(\gamma+\alpha) |\psi_{x}|^{2}\right]\\[0.2cm] |\psi_{x}|^{2}=\frac{\alpha_{S}}{\beta_{1}+\beta_{2}}(T_{c_{+}}-T)\\[0.2cm] |\psi_{y}|=0 \end{array} \right.\;\;for\;\;T_{c_{-}}<T\leq T_{c_{+}}\;, \end{equation} \begin{equation} \label{e08} \left\{ \begin{array}{l} M^{2}=\frac{1}{\beta_{M}}\left[\alpha_{M}(T_{N}-T^{\ast})-(\gamma+\alpha) |\psi_{x}|^{2}-\alpha|\psi_{y}|^{2}\right]\\ |\psi_{x}|^{2}=\frac{1}{\beta_{1}+\beta_{2}}\left[\alpha_{S}(T_{c_{+}}-T) -(\beta_{1}-\beta_{2})|\psi_{y}|^{2}\right]\\ |\psi_{y}|^{2}=\frac{\alpha_{S}}{\beta_{1}+\beta_{2}}(T_{c_{-}}-T) \end{array} \right.\;\;for\;\;T\leq T_{c_{-}}\;, \end{equation} where \begin{equation} \label{e09} T_{c_{+}}=T_{c}-\frac{\gamma+\alpha}{\alpha_{S}}M^{2}\;, \end{equation} \begin{equation} \label{e00} T_{c_{-}}=T_{c}-\frac{1}{\alpha_{S}}\left[\alpha M^{2}+(\beta_{1}-\beta_{2}) |\psi_{x}|^{2}\right]\;. \end{equation} $T_{c}$ is the superconducting transition temperature in a system without the magnetism. The complete solution to \mbox{Eqs.$(\ref{e04})-(\ref{e08})$} that is the explicit formulae for $T_{c_{-}}$ and $T_{c_{+}}$ are given in Appendix A \mbox{$(\ref{a01})-(\ref{a02})$}. The magnetic moment changes as \mbox{$M^{2}=\frac{\alpha_{M}}{\beta_{M}}(T_{N}-T)$} for temperatures higher than temperature $T^{\ast}$, then suddenly saturates at \mbox{$T^{\ast}\;\; (T^{\ast} \sim T_{c})$} and becomes constant below this temperature : \mbox{$M^{2}=\frac{\alpha_{M}}{\beta_{M}}(T_{N}-T^{\ast})$} in a normal (not superconducting) state. This temperature dependence of the magnetic moment is consistent with the measurements by Aeppli et al \cite{1,10}. They observed a kink at \mbox{$T^{\ast} \sim T_{c}$} and almost constant value of the magnetic Bragg intensity below \mbox{$T^{\ast}$} for magnetic field \mbox{$H>H_{c_{2}}$} that is when the system was not superconducting. The \mbox{$T^{\ast}$} temperature is introduced in our free energy \mbox{$(\ref{e2})$} rather artificially in order to fit the existing experimental data \cite{1,10}. We shall comment more on this issue further in the text.\\ \hspace*{0.5cm} From the free energy density $F_{S}\;\;(\ref{e3})$ we get the linear pressure dependence of the superconducting transition temperature: \begin{equation} \label{e5} T_{c}=T_{c}^{0}-a_{0}p\;, \end{equation} where $a_{0}$ is a constant coefficient and $T_{c}^{0}$ - a critical temperature $T_{c}$ at zero pressure $(p=0)$. We also assume the squared magnetic moment to be a linear pressure function: \begin{equation} \label{e6} M^{2}=M_{0}^{2}\frac{p_{N}-p}{p_{N}}\;, \end{equation} where $M_{0}$ is a magnetic moment at $p=0$, \mbox{$M_{0}=M(T,p=0)$} and $p_{N}$ $(\mbox{$p_{N}$}=\mbox{$p_{c}\simeq 4 kbar$})$ is a pressure at which the antiferromagnetism vanishes. In the superconducting system described by the free energy density $(\ref{e1})$ the magnetic and the superconducting terms compete in the coupling term $(\ref{e4})$. This interaction leads to the split of critical temperature \mbox{$T_{c}$} into \mbox{$T_{c_{-}}$} and \mbox{$T_{c_{+}}$} \cite{9} : \begin{equation} \label{e7} T_{c_{+}}-T_{c_{-}}=\frac{|\gamma|}{\alpha_{S}}\frac{\beta_{1}+\beta_{2}} {2\beta_{2}}M^{2}\;. \end{equation} \hspace{0.5cm} One can establish the pressure dependence of \mbox{$T_{c_{+}}$} and \mbox{$T_{c_{-}}$} from Eqs.$(\ref{e5})$ and $(\ref{e6})$ : \begin{eqnarray} \label{e8} T_{c_{+}} & = & T_{c_{+}}^{0}-a_{+}p\;,\\ \label{e9} T_{c_{-}} & = & T_{c_{-}}^{0}-a_{-}p\;, \end{eqnarray} where \begin{eqnarray} \label{e10} T_{c_{+}}^{0} & = & T_{c}^{0}+\frac{|\gamma|-\alpha}{\alpha_{S}}M_{0}^{2}\;,\\ \label{e11} T_{c_{-}}^{0} & = & T_{c}^{0}-\frac{1}{\alpha_{S}}\left(\alpha+ \frac{\beta_{1}-\beta_{2}}{2\beta_{2}}|\gamma|\right)M_{0}^{2}\;,\\ \label{e12} a_{+} & = & a_{0}+\left(\frac{|\gamma|-\alpha}{\alpha_{S}}\right) \frac{M_{0}^{2}}{p_{N}}\;,\\ \label{e13} a_{-} & = & a_{0}-\frac{1}{\alpha_{S}}\left(\alpha+\frac{\beta_{1} -\beta_{2}}{2\beta_{2}}|\gamma|\right)\frac{M_{0}^{2}}{p_{N}}\;. \end{eqnarray} \hspace{0.5cm} To obtain a proper pressure behavior (Fig.1) the following constraints must be fulfilled : \begin{equation} \label{e14} a_{+}>a_{0} \;\;\;\;\;{\rm and}\;\;\;\;\;a_{-}<a_{0}\;. \end{equation} Together with a condition $(\ref{e0})$ they give the relations between \\ the \mbox{Ginzburg-Landau} coefficients : \begin{equation} \label{e15} \frac{1}{2}\left(1-\frac{\beta_{1}}{\beta_{2}}\right)|\gamma|<\alpha< \frac{1}{4}\left(3-\frac{\beta_{1}}{\beta_{2}}\right)|\gamma|\;. \end{equation} \hspace{0.5cm} Now we turn to the magnetic Bragg scattering measurements \cite{1,10} (Fig. 2a). Since the neutron scattering intensity is proportional to $M^{2}$ we look at the magnetic moment and analyse it as a function of temperature. Taking into account that the coupling coefficients $\alpha$ and $\gamma$ $(\ref{e4})$ are expected to be much smaller than the other G-L coefficients \cite{14} and therefore neglecting higher than the linear in $\alpha$ and $\gamma$ terms from \mbox{Eqs.$(\ref{e1})-(\ref{e4})$} we obtain~: \begin{equation} \label{e16} M^{2}=M_{c}^{2}+a_{M}T\;, \end{equation} where \begin{equation} \label{e17} M_{c}^{2}=\frac{\alpha_{M}}{\beta_{M}}(T_{N}-T^{\ast})-a_{M}T_{c}\;, \end{equation} and \begin{eqnarray} \label{e18} a_{M}=\frac{\alpha_{S}}{\beta_{M}}\frac{\gamma+\alpha}{\beta_{1}+\beta_{2}} & for & T_{c_{-}}<T\leq T_{c_{+}}\;,\\ \label{e19} a_{M}=\frac{\alpha_{S}}{\beta_{M}}\frac{2\alpha+\gamma}{2\beta_{1}} & for & T \leq T_{c_{-}}\;. \end{eqnarray} In $M^{2}$ given by \mbox{Eqs.$(\ref{e16})-(\ref{e19})$} a discontinuity arises at \mbox{$T=T_{c_{-}}$} with a jump of the second order of magnitude in $\alpha$ and $\gamma$. Therefore it is negligible in the linear approximation. We present the full formula for $M^{2}$ in Appendix A \mbox{Eqs.$(\ref{a03})-(\ref{a07})$}. It can be shown that even within this general description the results of this section still hold.\\ There are two characteristic temperatures - $T_{c_{+}}$ and $T_{c_{-}}$ distinguished by the superconducting phase transitions, hence the change in the temperature dependence of the magnetic moment due to superconductivity can take place at one of these temperatures. For $M^{2}$ increasing with the temperature up to $T_{c_{+}}$ and decreasing then, that is for a kink at $T=T_{c_{+}}$ the condition : \begin{eqnarray} \label{e20} a_{M}>0 & for & T<T_{c_{+}} \end{eqnarray} is required, while for a kink at $T=T_{c_{-}}$ the following constraints are to be fulfilled: \begin{eqnarray} \label{e21} a_{M}<0 & for & T_{c_{-}}<T<T_{c_{+}} \end{eqnarray} and \begin{eqnarray} \label{e22} a_{M}>0 & for & T<T_{c_{-}}\;. \end{eqnarray} The condition $(\ref{e20})$ leads to the inequality: \begin{equation} \label{e23} \alpha>|\gamma|\;, \end{equation} whereas from $(\ref{e21})\;{\rm and}\;(\ref{e22})$ follows that: \begin{equation} \label{e24} \frac{1}{2}|\gamma|<\alpha<|\gamma|\;. \end{equation} It is evident that the condition $(\ref{e23})$ is inconsistent with the pressure relation $(\ref{e15})$, while the conditions $(\ref{e15})$ and $(\ref{e24})$ yield the relation \mbox{$\frac{\beta_{1}}{\beta_{2}}<1$} which is in contradiction with the specific heat measurements data \cite{15}. Put into words thermodynamics requires that if the magnetic moment is reduced when the sample becomes superconducting then the tendency to become superconducting will be increased if the magnetism is removed. This implies that the continuation of the phase line between normal and superconducting phases for $p>p_{c}$ should lie above $T_{c_{+}}$ if it is extrapolated back to low pressure in clear contrast to the data shown in Fig.1 and also more recent data of Boukhny et al. \cite{20}.\\ \hspace*{0.5cm} Therefore we conclude that it is {\bf not possible to explain the pressure and neutron scattering data in a frame of the free energy density ${\bf (\ref{e1})-(\ref{e4})}$} and the decrease in the magnetic Bragg intensity cannot be attributed to the decrease in $M$ only if it is assumed that $T_{c_{+}}-T_{c_{-}}$ is due to the coupling with magnetism. This paper does not address the alternative possibility that the splitting of $T_{c}$ is due instead to the coupling of the superconductivity to the charge density wave \cite{19,21} except to note that even if the effect of magnetism is only to reduce both $T_{c_{+}}$ and $T_{c_{-}}$ due to a pair breaking mechanism \cite{21} then there should still be a break in slope in $T_{c_{+}}$ at the pressure where magnetism is suppressed.\\ \hspace*{0.5cm} In the next paragraph we analyse the possibility of a rotation and decrease of the magnetic moment suggested by Blount et al. \cite{16} and Joynt \cite{17}. The rotation of magnetic moment can be equivalently described by an additional linearly independent magnetic moment $\bar{m}\;\;(\bar{m}\bot\bar{M})$ included. \section{\bf Two magnetic moment model} \hspace{0.5cm} In this section we consider the possibility that the magnetic moment rotates at the temperature of the order of $T_{c}$ in such a way that the observed Bragg scattering intensity is reduced. This requires two components of magnetisation. Therefore we propose a revised G-L free energy density: \begin{equation} \label{e25} F=F_{S}+F_{M}+F_{m}+F_{SM}+F_{sm}\;, \end{equation} where \begin{equation} \label{e026} F_{M}=\left\{ \begin{array}{l} \alpha_{M}\left(T^{\ast}-T_{N}\right)M^{2}+\frac{1}{2}\beta_{M}M^{4}\;\; for\;\;T\leq T_{m}\\ \alpha_{M}\left(T-T_{N}\right)M^{2}+\frac{1}{2}\beta_{M}M^{4}\;\; for\;\;T>T_{m}\;, \end{array} \right. \end{equation} \begin{equation} \label{e26} F_{m}=\alpha_{m}(T-T_{m})m^{2}+\frac{1}{2}\beta_{m}m^{4}\;, \end{equation} \begin{equation} \label{e27} F_{sm}=\gamma'|\bar{m}\bar{\psi}|^{2}+\alpha'm^{2}|\bar{\psi}|^{2}\;, \end{equation} and $F_{S},\;F_{M},\;F_{SM}$ are given by Eqs.$(\ref{e2})-(\ref{e4})$. $T_{m}$ is the N\'eel temperature of the magnetic moment $\bar{m}$ and $T_{m}\sim T_{c}$. The new coefficients $\alpha_{m}$ and $\beta_{m}$ in \mbox{$(\ref{e26})$} are positive. This free energy is correct to the fourth order in the space of $\bar{M},\;\bar{m}\;{\rm and}\;\bar{\psi}$. For the sake of simplicity we have neglected the coupling term between the two magnetic moments and the superconducting order parameter \mbox{$(\;mM(\psi_{x}\psi_{y}^{\ast}+\psi_{x}^{\ast}\psi_{y})\;)$} here, assuming it to have a little effect on the results. Another free energy term involving $\bar{M}$ and $\bar{m}$ \mbox{$(\;\sim m^{2}M^{2}\;)$} is included implicitly in $T_{m}$ and $T^{\ast}$ by a proper diagonalization of the magnetic part of the free energy \mbox{( Appendix B ).} As it is seen from $(\ref{e25})$, a magnetic moment $M$ is constant in the absence of superconductivity and equals: \begin{equation} \label{e28} M^{2}=\frac{\alpha_{M}}{\beta_{M}}(T_{N}-T^{\ast})\;. \end{equation} This approximation is correct for temperatures lower than a certain temperature of the order of $T_{c}$. We believe that this assumed temperature dependence of $M^{2}$ is due to a change in a Fermi surface and it is exclusively of the microscopic origin. However in the \mbox{Appendix B} we present a phenomenological explanation of this fact, when relation $(\ref{a4})$ is fulfilled. In this interpretation $M^{2}$ becomes constant below the temperature $T_{m}\;\;(\ref{e26},\ref{a7})$, that is the temperature at which the magnetic moment $\bar{m}$ appears. Although \mbox{$T_{m}\sim T_{c}$}, this reasoning is valid only if \mbox{$T_{m}>T_{c_{+}}$} which seems to be in agreement with the experimental data \cite{1,10}.\\ \hspace*{0.5cm} Proceeding in the same way as in section 2., from the pressure requirements \mbox{$(\ref{e0},\;\ref{e14})$} and the free energy density $(\ref{e25})$, we obtain the following conditions: \begin{eqnarray} \label{e29} (|\gamma|-\alpha)M_{0}^{2} & > & \alpha'm_{0}^{2}\;,\\ \label{e30} \left[\frac{1}{4}\left(3-\frac{\beta_{1}}{\beta_{2}}\right)|\gamma| -\alpha\right]M_{0}^{2} & > & \left[\alpha'+\frac{\beta_{1}+\beta_{2}} {4\beta_{2}}\gamma'\right]m_{0}^{2}\;, \end{eqnarray} where we have assumed, that $\bar{m}$ disappears at the same critical pressure $p_{N}$ as $\bar{M}$ does $(\ref{e6})$: \begin{equation} \label{e31} m^{2}=m_{0}^{2}\frac{p_{N}-p}{p_{N}} \end{equation} otherwise a kink in the pressure dependence of $T_{c_{-}}$ and $T_{c_{+}}$ should be observed, which is not the case (see Fig.1) \cite{2,7,8}.\\ \hspace*{0.5cm} Since there is no coupling terms between $\bar{m}$ and $\bar{M}$ in the free energy density $(\ref{e25})$, it yields the same temperature dependence of $M^{2}$ as in \mbox{Eqs.$(\ref{e16})-(\ref{e19})$.} Therefore in order to obtain an appropriate temperature behavior of $M^{2}$ \cite{1,10} (Fig. 2a) either $(\ref{e23})$ or $(\ref{e24})$ must be satisfied.\\ \hspace*{0.5cm} Now we are able to give the final conditions for the \mbox{G-L coefficients} in the free energy density which agrees with the experiments \cite{1,2,7,8,10} discussed in this paper. For $M^{2}$ increasing with the temperature up to $T=T_{c_{-}}$ and decreasing above this temperature the conditions $(\ref{e24})$ and \mbox{$(\ref{e29})-(\ref{e30})$} are to be held. They lead to a simple constraint on $\alpha'$, which is necessary but not sufficient: \begin{eqnarray} \label{e32} |\gamma|M_{0}^{2} & > & 2\alpha' m_{0}^{2}\;. \end{eqnarray} When $M^{2}$ as a function of temperature has a kink at \mbox{$T=T_{c_{+}}$}, that is increases below this temperature and decreases above it, the conditions $(\ref{e23})$ and \mbox{$(\ref{e29})-(\ref{e30})$} must be fulfilled and they yield the negative value of $\alpha'$ : \begin{eqnarray} \label{e33} \alpha' & < & 0 \;. \end{eqnarray} \section{\bf ${\bf \left(\frac{1}{2},0,1\right)}$ neutron scattering intensity} \hspace{0.5cm} We are going to consider both the magnetic moments $\bar{M}$ and $\bar{m}$ more thoroughly now. Here again we restrict the calculations to the linear in $\alpha$, $\beta$, $\alpha'$ and $\gamma'$ coupling coefficients terms, what yields a negligible in this approximation $M^{2}$ and $m^{2}$ discontinuity at $T_{c_{-}}$. A minimisation of the free energy \mbox{$(\ref{e25})$} as a magnetic moment $\bar{m}$ function leads to the temperature dependence of $m^{2}$ : \begin{equation} \label{e34} m^{2}=m_{c}^{2}+(a_{m}-\frac{\alpha_{m}}{\beta_{m}})T\;, \end{equation} where \begin{equation} \label{e35} m_{c}^{2}=\frac{\alpha_{m}}{\beta_{m}}T_{m}-a_{m}T_{c}\;, \end{equation} and \begin{eqnarray} \label{e36} a_{m}=\frac{\alpha_{S}}{\beta_{m}}\frac{\alpha'}{\beta_{1}+\beta_{2}} & for & T_{c_{-}}<T\leq T_{c_{+}}\;,\\ \label{e37} a_{m}=\frac{\alpha_{S}}{\beta_{m}}\frac{2\alpha'+\gamma'}{2\beta_{1}} & for & T\leq T_{c_{-}}\;. \end{eqnarray} We assume throughout this paper that the magnetic moments lie in the basal plane since the easy magnetic directions are confined to this plane. In the previous chapters we were considering the neutron reflections at the \mbox{reciprocal-lattice} point \mbox{$\bar{q_{1}}=(1,\frac{1}{2},0)$} \cite{1,10} (Fig. 2a). The magnetic Bragg scattering measurements revealed a different temperature dependence of the neutron scattering intensity at \mbox{$\bar{q_{2}}=(\frac{1}{2},0,1)$} \cite{18} (Fig. 2b). Below the temperature of the order of $T_{c}$ the \mbox{$(\frac{1}{2},0,1)$} intensity ceases to evolve and becomes constant. Actually, Aeppli et al. \cite{18} did not go with temperature low enough to be definitely positive about the $T$ independence of the measured intensity in the whole temperature range below $T_{c}$. Nevertheless, we assume here a constant value of \mbox{$(\frac{1}{2},0,1)$} neutron scattering intensity below $T_{c_{+}}$, that is we suggest this effect to be due to superconductivity. The neutron scattering intensity at the \mbox{reciprocal-lattice} point $\bar{q}$ reflects the magnetic vectors perpendicular to the $\bar{q}$ vector. For the sake of simplicity we choose a magnetic moment \begin{equation} \label{e40} \bar{M_{1}}=\bar{M}+\bar{m}\; \end{equation} perpendicular to \mbox{$\bar{q_{2}}=(\frac{1}{2},0,1)$} which means that $M_{1}^{2}$ is detected in \mbox{$(\frac{1}{2},0,1)$} measurements. On this particular magnetic orientation we want to check, without going into the detailed calculation of a general case, whether the two magnetic moment model can interpret both neutron scattering experiments. It will yield some additional constraints on the \mbox{G-L} free energy coefficients \mbox{$(\ref{e25})-(\ref{e27})$}. One of the possible considered configurations of the magnetic and reciprocal-lattice vectors, where instead of \mbox{$\bar{q_{1}}=(1,\frac{1}{2},0)$} and \mbox{$\bar{q_{2}}=(\frac{1}{2},0,1)$} their projections on the XY plane - \mbox{$(1,\frac{1}{2})$} and \mbox{$(\frac{1}{2},0)$} were plotted, is presented in Fig.3. $\bar{M}$ is the magnetic moment seen in $(1,\frac{1}{2},0)$ neutron scattering, while $\bar{M_{1}}$ is detected in $(\frac{1}{2},0,1)$ measurements. The temperature dependence of $M^{2}$ has been considered in the previous paragraphs of this paper \mbox{$(\ref{e16}-\ref{e19},\ref{e28})$}. According to \cite{18} $M_{1}^{2}$ is temperature independent for \mbox{$T\leq T_{c} \sim T_{c_{+}}$}: \begin{eqnarray} \label{e38} M_{1}^{2}=const & for & T<T_{c_{+}}\;. \end{eqnarray} Assuming the temperature dependent corrections to $M$ $\mbox{$(\ref{e16})$}- \mbox{$(\ref{e19})$}$ and $m$ $\mbox{$(\ref{e34})$}-\mbox{$(\ref{e37})$}$ to be small, experimentally estimated as about 5\% of the total magnetic moments values \cite{1,10}, we linearize $M$ and $m$ in $T$ and insert them into \mbox{Eq.$(\ref{e40})$}. Then the condition $(\ref{e38})$ leads to the following constraints on the \mbox{G-L} coefficients \begin{eqnarray} \label{e46} f(\frac{\gamma+\alpha}{\beta_{1}+\beta_{2}},\frac{\alpha'} {\beta_{1}+\beta_{2}})=0\;,\\ \label{e47} f(\frac{\gamma+2\alpha}{2\beta_{1}},\frac{\gamma'+2\alpha'} {2\beta_{1}})=0\;, \end{eqnarray} where \begin{equation} \label{e48} f(x,y)=\frac{\alpha_{S}}{\beta_{M}}x+\frac{1}{\beta_{m}} (\alpha_{S}y-\alpha_{m})\;. \end{equation} We solve the Eqs.$(\ref{e46})$ and $(\ref{e47})$ and \mbox{obtain :} \begin{equation} \label{e49} \beta_{1}=\frac{\alpha_{S}}{2\alpha_{m}}\left[\frac{\beta_{m}}{\beta_{M}} (\gamma+2\alpha)+\gamma'+2\alpha'\right]\;, \end{equation} \begin{equation} \label{e50} \beta_{2}=\frac{\alpha_{S}}{2\alpha_{m}}\left[\frac{\beta_{m}}{\beta_{M}} \gamma-\gamma'\right]\;. \end{equation} According to experiments \cite{15}, $\beta_{1}$ and $\beta_{2}$ coefficients should obey a following relation : \begin{equation} \label{e51} \beta_{1}>\beta_{2}>0\;. \end{equation} {}From \mbox{$(\ref{e49})-(\ref{e51})$} we have then : \begin{equation} \label{e52} \frac{\beta_{m}}{\beta_{M}}\gamma-\gamma'>0 \end{equation} and \begin{equation} \label{e53} \frac{\beta_{m}}{\beta_{M}}\alpha+\gamma'+\alpha'>0\;. \end{equation} Since $\gamma\;(\ref{e4})$ is negative, inequality $(\ref{e52})$ leads to a negative $\gamma'$ value and finally relation $(\ref{e52})$ is equivalent to : \begin{equation} \label{e54} \gamma'=-|\gamma'|\;,\;\;\;\;|\gamma'|>\frac{\beta_{m}}{\beta_{M}}|\gamma|\;. \end{equation} Therefore we have obtained conditions \mbox{$(\ref{e49})-(\ref{e50})$} and \mbox{$(\ref{e53})-(\ref{e54})$} which are to be fulfilled by \mbox{G-L} free energy coefficients. However we cannot forget about the constraints which follow from the $M^{2}$ temperature evolution requirements \mbox{$(\ref{e23})-(\ref{e24})$} and these which are necessary to fit the pressure data \mbox{$(\ref{e29})-(\ref{e30})$}. One can check easily that the conditions $(\ref{e23})$ ( kink in $M^{2}$ at $T=T_{c_{+}}$ ) and \mbox{$(\ref{e53})-(\ref{e54})$} lead to a negative value of $\alpha'$, while the constraint $(\ref{e24})$ ( kink in $M^{2}$ at \mbox{$T=T_{c_{-}}$ )} along with \mbox{Eqs.$(\ref{e53})-(\ref{e54})$} yield a positive $\alpha'$ value. From \mbox{Eqs.$(\ref{e29})-(\ref{e30})$} we get more information about the magnetic moments values at pressure $p=0$, that is $M_{0}\; (\ref{e6})$ and $m_{0}\;(\ref{e31})$. It is more convenient for this purpose to use the experimentally established $\frac{\beta_{2}}{\beta_{1}}$ ratio : \mbox{$\frac{\beta_{2}}{\beta_{1}}\simeq 0.4$} \cite{15}, just to get rid of $\beta_{1}$ and $\beta_{2}$ coefficients in $(\ref{e30})$. The relation \mbox{$\frac{\beta_{2}}{\beta_{1}}=0.4$} along with \mbox{$\beta_{1}\; (\ref{e49})$} and \mbox{$\beta_{2}\;(\ref{e50})$} formulae allow the reduction of one of the coupling coefficients through the equation : \begin{equation} \label{e55} \alpha'=\frac{7}{4}(|\gamma'|-\frac{\beta_{m}}{\beta_{M}}|\gamma|) +\frac{\beta_{m}}{\beta_{M}}(|\gamma|-\alpha)\;, \end{equation} so we can consider $\gamma',\;\gamma$ and $\alpha$ parameters as the only independent in all the conditions. It is straightforward to show that $\alpha'$ given by \mbox{Eq.$(\ref{e55})$} obeys the \mbox{Eqs.$(\ref{e23})-(\ref{e24})$} and \mbox{$(\ref{e53})-(\ref{e54})$}. Returning to $M_{0}$ and $m_{0}$ magnitudes, for \mbox{$\alpha>|\gamma|\; (\ref{e23})$}, we obtain from \mbox{$(\ref{e29})-(\ref{e30})$}, that \begin{equation} \label{e56} m_{0}^{2}>g_{0}M_{0}^{2}\;, \end{equation} where \begin{displaymath} g_{0}=max\left\{\frac{\alpha-|\gamma|}{|\alpha'|}, \frac{\alpha-\frac{1}{8}|\gamma|}{|\alpha'|+\frac{7}{8}|\gamma'|}\right\}\;. \end{displaymath} The condition above should be fulfilled when a kink in \mbox{$\left(1,\frac{1}{2},0\right)$} neutron scattering intensity appeares at $T_{c_{+}}\;(\ref{e23})$. In order to have $m_{0},M_{0}$ solutions of \mbox{$(\ref{e29})-(\ref{e30})$} when condition $(\ref{e24})$ is held, that is in a case of the \mbox{$\left(1,\frac{1}{2},0\right)$} neutron scattering peak at $T_{c_{-}}$, another constraint is to be fulfilled : \begin{equation} \label{e57} \alpha'<\frac{7}{8}|\gamma'|\;. \end{equation} Inequality $(\ref{e57})$ is a necessary condition to make sense to the relations $(\ref{e29})$ and $(\ref{e30})$.\\ Finally, we obtain from \mbox{$(\ref{e29})-(\ref{e30})$} the constraint on the relative $m_{0}$ and $M_{0}$ values : \begin{equation} \label{e58} \frac{\alpha-\frac{1}{8}|\gamma|}{\frac{7}{8}|\gamma'|-\alpha'}< \frac{m_{0}^{2}}{M_{0}^{2}}<\frac{|\gamma|-\alpha}{\alpha'}\;, \end{equation} and another condition which follows straightly from $(\ref{e58})$ : \begin{equation} \label{e59} |\gamma\gamma'|-|\gamma|\alpha'-|\gamma'|\alpha>0\;. \end{equation} We have been looking here at the additional constraints on the fourth order coefficients in the \mbox{Ginzburg-Landau} free energy, that follow from the requirement of a constant magnetic moment detected in \mbox{$\left(\frac{1}{2},0,1\right)$} neutron scattering measurements \cite{18} (Fig. 2b). We have assumed $T_{c_{+}}$ as a characteristic temperature at which the magnetic moment $M_{1}\;(\ref{e38})$ becomes constant. Nevertheless it is straightforward to show that $M_{1}$ cannot be constant above $T_{c_{+}}$. Let us look at the temperatures \mbox{$T>T_{m}$} first. Since $T_{m}$ is the N\'eel temperature for $\bar{m}\;(\ref{a6})$, there is only one magnetic moment $\bar{M}$ left at $T>T_{m}$. $M_{1}$ is simply $M$'s projection on a particular direction ( Fig.3 ) and shows the same temperature dependence as $M$ does $(\ref{a9})$. Therefore $M_{1}$ is a decreasing function of temperature for \mbox{$T>T_{m}$} as $M$ is $(\ref{a9})$. In the temperature range \mbox{$T_{c_{+}}<T<T_{m}$}, on the other hand, we obtain from the free energy $(\ref{a8})$ a constant $M^{2}$ value $(\ref{e28})$ and \mbox{$m^{2}=\frac{\alpha_{m}}{\beta_{m}}\left(T_{m}-T\right)$}. Therefore $(\ref{e40})$ cannot lead to a constant $M_{1}$ value, otherwise $\alpha_{m}=0$ and $\bar{m}$ vanishes, what makes no sense for this approach. \section{\bf Conclusions} \hspace{0.5cm} We have considered superconducting \mbox{$U\!Pt_{3}$} in zero magnetic field. Our interest has been focused on the hydrodynamic pressure \cite{2,7,8} and neutron scattering experiments \cite{1,10,18}. We have shown that the pressure dependence of the transition temperatures and the abrupt change in the \mbox{$\left(1,\frac{1}{2},0\right)$} neutron scattering intensity at \mbox{$T\sim T_{c}$} \cite{1,10} cannot be explained quantitatively within a simple two component superconducting order parameter which couples to one component antiferromagnetism. As one way of reconciling this problem we have suggested the existence of another magnetic moment which emerges at \mbox{$T\sim T_{c}$}. This generalized approach of the two independent magnetic moments coupling to the superconductivity allowed us to obtain a concise picture of discussed phenomena and yields several stringent constraints on the fourth order coefficients in the \mbox{Ginzburg-Landau} free energy density $(\ref{e25})$. We have concluded that the kink in a \mbox{$\left(1,\frac{1}{2},0\right)$} neutron scattering intensity may exist at $T_{c_{+}}$ when $(\ref{e23})$ and \mbox{$(\ref{e29})-(\ref{e30})$} relations between the \mbox{G-L} coefficients are obeyed or at $T_{c_{-}}$ under the condition of $(\ref{e24})$ and \mbox{$(\ref{e29})-(\ref{e30})$}. If we interpret the results of \mbox{$\left(\frac{1}{2},0,1\right)$} Bragg magnetic scattering experiments \cite{18} as characteristic feature for all temperatures below $T_{c}$ and assume the magnetic moments orientation as in Fig. 3, we can express $\beta_{1}$ and $\beta_{2}$ \mbox{G-L} coefficients in terms of the coupling constants \mbox{$(\ref{e49})-(\ref{e50})$}. The requirement \mbox{$\beta_{1}>\beta_{2}>0$} leads to a negative value of a coupling constant $\gamma'\;(\ref{e27},\ref{e54})$ and negative $\alpha'\; (\ref{e27})$ coefficient value when a peak in \mbox{$\left(1,\frac{1}{2},0\right)$} neutron scattering intensity is at $T_{c_{+}}$ or positive $\alpha'$ value for a peak at $T_{c_{-}}$. These considerations yield also some constraints on the zero pressure magnetic moments values \mbox{$(\ref{e56},\ref{e58})$} and coupling coefficients \mbox{$(\ref{e57},\ref{e59})$}. We have evaluated \mbox{$(\ref{e56})-(\ref{e59})$} constraints for the experimentally established ratio \mbox{$\frac{\beta_{2}}{\beta_{1}}\simeq 0.4$} \cite{15}. This given value of $\frac{\beta_{2}}{\beta_{1}}$ allows us to express one of the G~-~L coupling coefficients in terms of the others $(\ref{e55})$.\\ We have considered two magnetic moments in a crystal basal plane only. However, we cannot exclude any of them out of this plane. There is always a possibility of a magnetic structure following a recently discovered structural modulation in a crystal \cite{19}. Unfortunately the resolution of a neutron scattering measurements may be to small to be decisive. For the completness of the picture it should be added that despite a large number of experimental evidence the main facts seems to be unsettled. It concernes the phase diagram in the p-T plane measured by Boukhny et al. \cite{20} where the slope of $T_{c_{-}}$ curve is positive and the condition $(\ref{e0})$ does not hold. Moreover the recent x-ray resonant magnetic and neutron magnetic scattering measurements \cite{21} show no correlation between the split superconducting transition and the weak antiferromagnetic order in $U\!Pt_{3}$ and as they also find no evidence of magnetic moment rotation their results together with the conclusions of the paper suggest other possible issues like symmetry-breaking fields of structural origin \cite{19} or the existence of two one dimensional superconducting states. \section*{\bf Acknowledgements} We would like to thank G. Aeppli for fruitful comments and providing us with his recent experimental results as well as R. Joynt and L. Taillefer for helpful letters.\\ This work was supported by the European Community's Action for Cooperation in Sciences and Technology with Central and Eastern European Countries Fellowship \mbox{ contract $n^{\circ}:\;ERB-CIPA-92-2218$.} \section*{\bf Appendix A} \begin{equation} \label{a01} T_{c_{+}}=T_{c}-\frac{\alpha+\gamma}{\alpha_{S}}\frac{\alpha_{M}} {\beta_{M}}\left(T_{N}-T^{\ast}\right)\;, \end{equation} \begin{equation} \label{a02} T_{c_{-}}=T_{c}+\frac{\alpha_{M}}{\alpha_{S}} \frac{\gamma\left(\beta_{1}-\beta_{2}\right)-2\beta_{2}\alpha} {2\beta_{2}\beta_{M}-\gamma\left(\alpha+\gamma\right)}\left(T_{N}-T^{\ast} \right)\;, \end{equation} \begin{equation} \label{a03} M^{2}=M_{c}^{2}+a_{M}T\;, \end{equation} \begin{equation} \label{a04} M_{c}^{2}=\left\{ \begin{array}{lll} \frac{\beta_{1}+\beta_{2}}{\lambda_{+}} \left[\alpha_{M}\left(T_{N}-T^{\ast}\right)-\frac{\alpha_{S}} {\beta_{1}+\beta_{2}}\left(\alpha+\gamma\right)T_{c}\right] & for & T_{c_{-}}<T\leq T_{c_{+}}\\ \frac{2\beta_{2}}{\lambda_{-}}\left[2\alpha_{M}\beta_{1} \left(T_{N}-T^{\ast}\right)-\alpha_{S}\left(2\alpha+\gamma\right)T_{c}\right] & for & T\leq T_{c_{-}}\;, \end{array} \right. \end{equation} \begin{equation} \label{a05} a_{M}=\left\{ \begin{array}{lll} \frac{\alpha_{S}}{\lambda_{+}}\left(\alpha+\gamma\right) & for & T_{c_{-}}<T\leq T_{c_{+}}\\ \frac{2\alpha_{S}}{\lambda_{-}}\beta_{2}\left(2\alpha+\gamma\right) & for & T\leq T_{c_{-}}\;, \end{array} \right. \end{equation} \begin{equation} \label{a06} \lambda_{+}=\beta_{M}\left(\beta_{1}+\beta_{2}\right)- \left(\alpha+\gamma\right)^{2}\;, \end{equation} \begin{equation} \label{a07} \lambda_{-}=4\beta_{1}\beta_{2}\beta_{M}-4\beta_{2}\alpha \left(\alpha+\gamma\right)-\gamma^{2}\left(\beta_{1}+\beta_{2}\right)\;. \end{equation} \section*{\bf Appendix B} \hspace{0.5cm}The complete magnetic free energy for the magnetic moments $\bar{M}$ and $\bar{m}$ \mbox{$(\bar{m}\perp\bar{M})$} at the temperatures \mbox{$T<T^{\ast}$} is : \begin{equation} \label{a1} F_{magn}=A_{M}\left(T-T_{N}\right)M^{2}+\frac{1}{2}B_{M}M^{4}+ A_{m}\left(T-T^{\ast}\right)m^{2}+\frac{1}{2}B_{m}m^{4}+Cm^{2}M^{2}\;, \end{equation} where $T_{N}$ and $T^{\ast}$ are the N\'eel temperatures for $\bar{M}$ and $\bar{m}$ magnetic moments appropriately.\\ We assume $T_{N}>T^{\ast}$. From the minimisation of $F_{magn}$ one gets : \begin{equation} \label{a2} M^{2}=\frac{B_{M}B_{m}}{B_{M}B_{m}-C^{2}}\left[\frac{A_{M}}{B_{M}}T_{N} -\frac{A_{m}}{B_{m}}\frac{C}{B_{M}}T^{\ast}-\left(\frac{A_{M}}{B_{M}}- \frac{A_{m}}{B_{m}}\frac{C}{B_{M}}\right)T\right]\;, \end{equation} and \begin{equation} \label{a3} m^{2}=\frac{1}{B_{M}B_{m}-C^{2}}\left(A_{m}B_{M}T^{\ast}-A_{M}CT_{N}\right) -\frac{A_{m}}{B_{m}}T\;. \end{equation} For a particular choice of the coupling coefficient \begin{equation} \label{a4} C=\frac{A_{M}}{A_{m}}B_{m}\;, \end{equation} $M^{2}$ attains a constant value : \begin{equation} \label{a5} M^{2}=\frac{A_{M}A_{m}^{2}}{A_{m}^{2}B_{M}-A_{M}^{2}B_{m}}\left(T_{N}-T^{\ast} \right)\;, \end{equation} and \begin{equation} \label{a6} m^{2}=\frac{A_{m}}{B_{m}}\left(T_{m}-T\right)\;, \end{equation} where \begin{equation} \label{a7} T_{m}=\frac{A_{m}^{2}B_{M}-A_{M}^{2}B_{m}\frac{T_{N}}{T^{\ast}}} {A_{m}^{2}B_{M}-A_{M}^{2}B_{m}}T^{\ast}\;. \end{equation} The temperature $T^{\ast}$ should be of the order of $T_{N}$ to give a positive value of $T_{m}$. From $(\ref{a5})$ and $(\ref{a6})$ we can see that the magnetic free energy can be written as : \begin{equation} \label{a8} F_{magn}=\alpha_{M}\left(T^{\ast}-T_{N}\right)M^{2}+\frac{1}{2}\beta_{M}M^{4} +\alpha_{m}\left(T-T_{m}\right)m^{2}+\frac{1}{2}\beta_{m}m^{4}\;, \end{equation} for $T<T_{m}$, and \begin{equation} \label{a9} F_{magn}=\alpha_{M}\left(T-T_{N}\right)M^{2}+\frac{1}{2}\beta_{M}M^{4}\;, \end{equation} for $T>T_{m}$.\\ This is the free energy of two magnetic moments $(\ref{e026})-(\ref{e26})$ we use in this paper. The new G-L coefficients are given by the old \mbox{ones :} \begin{eqnarray} \label{a10} \alpha_{M}=A_{M}A_{m}^{2}\;,\\ \label{a11} \beta_{M}=A_{m}^{2}B_{M}-A_{M}^{2}B_{m}\;,\\ \label{a12} \alpha_{m}=A_{m}\;,\\ \label{a13} \beta_{m}=B_{m}\;. \end{eqnarray} These considerations are relevant only when $(\ref{a4})$ condition is fulfilled.